Regan Northwestern 0163D 15410

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 187

NORTHWESTERN UNIVERSITY

Quantized Vortices in Superfluid 3 He

A DISSERTATION

SUBMITTED TO THE GRADUATE SCHOOL

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

for the degree

DOCTOR OF PHILOSOPHY

Field of Theoretical Physics

By

Robert C. Regan

EVANSTON, ILLINOIS

December 2020
2

© Copyright by Robert C. Regan 2020

All Rights Reserved


3

ABSTRACT

Quantized Vortices in Superfluid 3 He

Robert C. Regan

Liquid 3 He is the paradigm system for studying exotic quantum mechanical phenomena

observed in nature. The normal state of 3 He is a strongly-correlated Fermi liquid with

strong ferromagnetic spin fluctuations. However, as we cool 3 He down to T = 0K, we

find something fascinating and unexpected. This exotic fluid remains a fluid even down

to absolute zero temperature, thus being called a ‘superfluid’. The ferromagnetic spin

fluctuations provide the pairing glue to bind spin-triplet Cooper pairs which form this

superfluid state. The superfluid phases of 3 He host topological defects known as quantized

vortices.

Quantized vortices are exotic topological phases of matter that are formed in the

superfluid phases of 3 He. Vortices are typically formed by rotation of the superfluid,

however they have also been observed to be generated even under zero rotation. Di↵erent

quantized vortices live in di↵erent superfluid phases of 3 He, and even within a single

superfluid phase; there can be multiple di↵erent vortex phases of matter. Understanding
4

the energetics and structures of quantized vortices in the superfluid phases is an important

theoretical task.

Symmetry breaking and phase transitions are another fundamental theme in condensed

matter physics. The quantized vortex can have a variety of di↵erent kinds of symmetries

and structures, they are connected to one another by phase transitions. Superfluid 3 He

is again the ideal candidate to exhibit and look for such phenomena. The vortex phases

within the bulk phases of superfluid 3 He have di↵erent symmetries and are connected

by first-order phase transitions. The vortex phases of the confined phases of superfluid
3
He have di↵erent symmetries but are now connected by second-order phase transitions.

Equilibrium and metastable phase transitions between vortices is driven by symmetry-

breaking and energetics.

In this thesis, I calculate the structure of vortices and vortex phase diagrams in mul-

tiple phases of superfluid 3 He. I first investigate the stability of singly-quantized vortices

in the B phase of 3 He, and calculate the phase diagram to determine the regions of sta-

bility for both equilibrium and metastable vortex phases in rotating superfluid 3 He-B. I

then stabilize confined phases of superfluid 3 He in Nafen, and investigate half-quantum

vortices within these phases of matter. I calculate the regions of energetic stability for the

half-quantized vortices and determine the vortex phase diagram in these confined phases.

Lastly, using the quasiclassical Eilenberger theory, I calculate the structure and excitation

spectrum of half-quantum vortices in confined superfluid 3 He.


5

Acknowledgements

I thank my advisor Jim Sauls for teaching me about life, physics and guiding me

through all these years. I also appreciate the level of rigor and discipline that I have

learned from working with Jim. It’s difficult to thank Jim in a few sentences all he

has done for me as Jim has helped me grow as a person deeply, not just a physicist. I

remember when I first met Jim, and I knew instantly from our real conservation that he

was the Professor I want to work with for my Ph.D. The way Jim does business is quite

impressive, and this is the style I will share with my future Ph.D students down the road.

A lot of my time at Northwestern was spent being a teaching assistant. I am very

thankful to have worked with Bill Halperin for many years. I learned a lot from Bill’s

style of teaching and his unique way of creating problems. I also appreciate all of the

personal help from Bill, and the endless discussions over the years. Whenever I needed

anything, I always knew I could rely on Bill and that’s very meaningful for me. I also

thank Bill for his unique deep patience, which has a↵ected me and my attitude on how I

handle certain situations. Lastly, I thank Bill for willing to be on my Ph.D committee.

I would like to thank Jens Koch for being on my Ph.D committee as well. I started out

taking classes with Jens my first year here, and I remember the style of his teaching that

a↵ected me a bit. His lectures, notes and discussions were always extremely organized and

well planned. Everything was properly referenced and thoroughly proof-read. I realize

many years later now how important this attitude is for researchers in physics. This results
6

in doing careful, meticulous, organized research leading to well written publications which

helps advance the field of physics.

I would like to thank the members of my research group, Hao Wu, Josh Wiman,

Oleksii Shevtsov, Wave Ngampruetikorn, Wei-Ting Lin, Mehdi Zarea and Junguang He.

I learned a lot from Hao early on in my Ph.D and enjoy our friendship very much still to

this day. I thank Oleksii for teaching me how to implement parallel computing techniques.

There is one member of my research group, Wei-Ting Lin, that I want to personally

acknowledge. Wei-Ting has been my best friend at Northwestern for years and helped

me with all of my research projects here. I learned how to write numerical codes and

many numerical strategies from Wei-Ting. I also appreciate the many weekends we spent

discussing theoretical problems together. It was also quite a journey making it to the

QFS conference in Canada bro! I’m not sure if I would have finished my Ph.D as quickly

if it wasn’t for Wei-Ting’s guidance and help.

I’d like to thank the members of Bill Halperin’s group, Yizhou Xin and Keenan Avers.

I spent my early years here discussing with Yizhou enjoying great friendship together.

I remember being quite blown away at Yizhou’s work ethic, and how he’d always be

working in the lab morning, night, during the night, anytime. This rubbed o↵ on me

a bit and motivated me to work harder as well. I also thank Keenan Avers for many

technical discussions on vortex physics and friendship over the years. It has also been of

great pleasure getting to know Keenan, and enjoying lunch/vortex discussions together.

Keenan has helped me a lot on my research projects and has always been there to give

his expertise. Despite being an experimentalist, Keenan has an unusually strong ability

to discuss any problem theoretically and this still impresses me very much.
7

I want to also thank a member of Jens Koch’s group, Ziwen Huang. I had the pleasure

of getting to know Ziwen over the years and developing a strong friendship. Ziwen works

all day, all night, seven days a week. I’ve always been impressed at his extreme work

ethic and this also motivated me to work harder. I’ve also learned a lot from Ziwen about

many things outside of physics. I also enjoyed playing basketball together, and his strong

competitive attitude. I hope one day that we can merge into a similar area in condensed

matter physics and work on a project together.

There are two physicists from Japan that have helped me and a↵ected my life very

much. First, I’d like to thank Yusuke Masaki. I met Yusuke in Tokyo at QFS and immedi-

ately we connected. Yusuke really showed me a di↵erent side of Japan, and our discussion

over nikuzushi changed my future forever. The realness, and the intense work ethic of

Yusuke is quite profound and I look forward to doing research in Japan together. We can

always find something to discuss whether it’s exotic Japanese food, Japanese culture, vor-

tex physics or other wild discussions, everything always comes natural together. Another

physicist I’ve been fortunate of meeting my last year at Northwestern is Hikaru Ueki. Our

week at Northwestern together was a very special week, and I had a real blast getting to

hang out so much and becoming best friends so naturally. Hikaru has taught me a lot

of physics already from our weekly intense discussions together. We both prefer to have

heavy discussions about physics, and seafood; and it’s this style that I appreciate very

much. I appreciate all the help that Yusuke and Hikaru have done in helping me pursue

Japan for my future physics career, and also in teaching me the Japanese language.

Lastly, I want to thank my parents and brothers. My parents have supported me

throughout my extreme ups and downs, and I would not be at Northwestern today if it
8

wasn’t for their intense continuous support in all ways. My brothers, well, it’s a bit hard

to describe my thanks to them in words. That’s not really how we do business together

either, my bros are my bros and we’ve done just about anything and everything together

my entire life.
9

Preface

In this thesis, I will discuss quantized vortices in the superfluid phases of 3 He. I

calculate vortex structures and vortex phase diagrams in multiple phases of superfluid
3
He as a function of temperature, pressure, and magnetic field. I first investigate the

singly-quantized vortex structures in the bulk B phase and proceed to calculating the

vortex phase diagram using strong-coupling Ginzburg-Landau theory. I then calculate

both the equilibrium and metastable phase transitions between di↵erent vortex states.

The vortex phase diagram is in excellent agreement with the experimentally reported

phase transitions in superfluid 3 He-B.

I then consider half-quantum vortices (HQV) in a confined nematic 3 He superfluid,

which is stabilized by infusing liquid 3 He into a low-density, highly anisotropic porous

aerogel called Nafen. Using strong-coupling Ginzburg-Landau theory, I identify two dif-

ferent ground state equilibrium vortex phases within the polar and polar-distorted chiral

A phase. In the polar phase, a HQV pair with a pure polar core is discovered. I then

lower the temperature and stabilize the polar-distorted chiral A phase which hosts a newly

discovered polar-distorted chiral HQV. The confined superfluid phase diagram is then cal-

culated in an axial magnetic field along the rotation axis, H = 0 370G ⌦, and I obtain

excellent agreement with the experimentally reported phase transitions observed in the

nematic Nafen phases of 3 He.


10

I lastly calculate the fermionic spectrum of an impurity lattice and polar half-quantum

vortices in the confined nematic polar phase of 3 He using the quasiclassical Eilenberger

equation in weak-coupling. The bound-state spectrum is understood by investigating the

local spectral density of states and spin-resolved density of states.


11

Table of Contents

ABSTRACT 3

Acknowledgements 5

Preface 9

Table of Contents 11

List of Tables 14

List of Figures 15

Chapter 1. Introduction 25

Chapter 2. Introduction to Quasiclassical Theory 30

2.1. Gorkov’s Equation 31

2.2. Quasiclassical Green’s Function 34

2.3. Eilenberger Transport Equations 38

2.4. Spectral Density and Currents 39

2.5. Riccati Transformation 41

2.6. P-Wave, Spin-Triplet Superfluid 3 He 43

2.7. Ginzburg-Landau Expansion 45

Chapter 3. The Vortex Phase Diagram of Rotating Superfluid 3 He-B 52


12

3.1. Introduction 53

3.2. Ginzburg-Landau Theory 58

3.3. Vortex States in Superfluid 3 He-B 67

3.4. Magnetic Susceptibility 86

3.5. Equilibrium & Metastability Transitions 90

3.6. Summary and Outlook 94

3.7. Acknowledgements 96

3.8. Appendix: Material Parameters 96

3.9. Appendix: Numerical Methods 97

3.10. Appendix: Benchmarking the GL Solver 101

Chapter 4. Half-Quantum Vortices in Superfluid 3 He Confined in Nafen 103

4.1. Ginzburg-Landau Theory 108

4.2. Order Parameter of a Half-Quantum Vortex 109

4.3. Numerical Solution of a Single Polar Half-Quantum Vortex 111

4.4. Analytical Solution of a Single Polar Half-Quantum Vortex 115

4.5. Pair of Polar Half-Quantum Vortices 122

4.6. Conversion of Half-Quantum Vortices into Singly-Quantized Vortices 130

4.7. Magnetic Field Evolution of a Polar Half-Quantum Vortex 135

4.8. Chiral Phase Order Parameter 137

4.9. Pair of Polar-distorted Chiral Half-Quantum Vortices 140

4.10. Superfluid Vortex Phase Diagram of 3 He Confined in Nafen 145

4.11. Conclusions 156


13

Chapter 5. Quasiparticle Bound States in Polar Half-Quantum Vortices 158

5.1. Local Density of States 158

5.2. Spin-Resolved Density of States 160

5.3. Conclusion 162

Chapter 6. Conclusion 163

References 166
14

List of Tables

3.1 Coefficients of a polynomial fit to the strong-coupling parameters


P (i)
from Ref. [1] of the form isc = n an pn . 96

3.2 Material parameters for 3 He vs. pressure, with the particle density

n = kf3 /3⇡ 2 from Ref. [2], the e↵ective mass, m⇤ , and Tc from Ref. [3],

the exchange interaction, F0a , is from Ref. [4], the Fermi velocity,

vf = ~kf /m⇤ , calculated from the Fermi wavelength, kf , and the

coherence length is ⇠0 = ~vf /2⇡ kB Tc . The strong-coupling parameters,


sc wc
i , in units of | 1 |, are from Ref. [1]. 97
15

List of Figures

3.1 The vortex core transition line, TV (p, H), for H = 0 G (solid green)

separating the A-core and D-core vortex phases of 3 He-B terminates

at a triple point (pvc , Tvc ) = (18.40 bar, 2.19 mK). In a magnetic field

the vortex core transition line extends to low temperatures down to

p = 0 bar in a magnetic field, as shown for H = 284 G (dashed green).

For comparison the Bulk AB transition lines for are shown in blue for

H = 0 G (solid) and H = 284 G (dashed). The A-core phase supercools

down to the metastability limit, TV⇤ (p, H), shown as the purple dashed

line for H = 284 G. Experimental data for the transition on cooling

(red diamonds) agrees well with the supercooling transition, while

the data point at p = 29.3 bar taken on warming (red square/circle)

agrees well with the calculated equilibrium vortex phase transition. The

experimental data is from Ref. [5]. 56

3.2 Left panel: The axially symmetric B-phase vortex (“o-vortex”) has a

hard core with a node in (r). Center panel: The axially symmetric

A-core vortex (“v-vortex”) has a suppressed, but non-vanishing,

condensate density in the core which is predominantly the order

parameter for the bulk A phase. Right panel: The D-core vortex has
16

a “double core” structure that spontaneously breaks axial rotation

symmetry. All plots are of the condensate density, | (r)|2 ⌘ Tr AA† ,

in units of that for the bulk B-phase, 2


B ⌘ Tr AB A†B . The solutions

of the GL equations for the o-vortex, A-core and D-core vortices

correspond to p = 10 bar and T = 0.25Tc , p = 34 bar and T = 0.75Tc ,

p = 20 bar and T = 0.55Tc , respectively. 68

3.3 Amplitudes and phases of the components of the o-vortex at P = 10

bar and T = 0.25 Tc , shown on a square grid with x and y ranging

from [ 5 ⇠, +5 ⇠]. The computational grid was 60 ⇠ ⇥ 60 ⇠ with grid

spacing h = 0.1⇠. The o-vortex retains the maximal symmetry of the

stationary vortex solutions for the B-phase; axial rotation symmetry

and time-reversal symmetry are preserved. Amplitudes with N = 0

and N = 2 vanish by symmetry. Thus, the o-vortex has vanishing

condensate density in the core. 72

3.4 Amplitudes and phases of the components of the A-core vortex at

P = 34 bar and T = 0.75 Tc , shown and computed on the same grid

as that in Fig. 3.3. The key amplitudes defining the A-core vortex are

the amplitudes with zero phase winding - the A phase C0+ and the

spin-polarized phase, C+0 . The amplitudes with winding number

N = 2, C 0 and C0 are also important in terms of the relative stability

between the A-core and D-core vortex states, as discussed in Sec. 3.3.3. 76

3.5 Amplitudes and phases of the components of the D-core vortex at

P = 20 bar and T = 0.55 Tc , shown and computed on the same


17

grid as that in Fig. 3.3. The key amplitudes and phases defining the

D-core vortex are those with N = 2 phase winding, C0 and C 0 . The

D-core accomodates the double phase winding by dissociation into two

N = 1 vortices, allowing the corresponding amplitudes to grow. This is

important for the stability of the D-core vortex relative to the A-core

and o-vortex, as discussed in Sec. 3.3.3. The dissociation of the N = 2

vortices is responsible for the broken axial symmetry that is clearly

shown in all the amplitudes and phases. 78

3.6 Left: Axially symmetric current density of the A-core vortex for p = 34

bar and T = 1.86 mK. The current is strongly suppressed to zero by

the growth of the A- and phases in the core for |r| . 2.5⇠. Center:

Anisotropic mass current flow field of the D-core vortex at P = 20bar,

T = 1.23 mK. Right: Expanded view of the current near the center of

the D-core showing the double vortex structure as the source of the

anisotropic current density. Currents are scaled in units of jc defined in

Eq. 3.28. 79

3.7 Axial mass current of the D-core vortex at p = 20 bar and T = 0.55 Tc .

The current density is scaled in units of jc . 83

3.8 Magnetization profiles, mA and mD , for the A- and D-core vortices,

respectively. Insets: density plots of the same. For the A-core phase:

p = 34 bar T = 0.75 Tc = 1.86 mK. For the D-core phase: p = 20 bar

T = 0.55 Tc = 1.23 mK. 87


18

A D
3.9 Susceptibility profiles, zz and zz , for the A- and D-core vortices,

respectively. Calculations are for the zero-field order parameters at


A
p = 34 bar just to the right and left of the zero-field transition line: zz

D
( zz ) is evaluated at T = 1.80 mK (T = 1.79 mK). Insets: density plots

of the same. 88

3.10 Field evolution of the equilibrium vortex transition temperature at

p = 34 bar. 91

3.11 Rate of convergence of the iterative solution of the GL equations

expressed in terms of the Free energy at each iteration, k, for the

o-vortex at P = 10 bar, T = 0.25 Tc normalized by the bulk B phase

free energy integrated over the same volume. The L-BFGS algorithm is

shown in green, while the relaxation algorithm is shown in red. 98

3.12 Free energies for the A-core and D-core vortex states versus magnetic

field calculated using our GL solver for the set I GL parameters at

p = 34 bar. The core contribution to the free energies are calculated as

in Ref.[6] by subtracting the bulk and hydrodynamic contributions to

the B phase energy, and normalizing the result in units of the bulk B

phase energy at each field. 102

P
4.1 Cooper pairing density, ↵,i |A↵i |2 of a single HQV in the polar phase

at p = 15bar, T = 0.95Tc . It is clear the vortex is superfluid in its core. 112


19

4.2 Amplitude and phase of a single HQV in the polar phase at T = 0.95Tc ,

p = 15bar. It can be described a singly-quantized vortex in the spin-up

condensate and a spin-down condensate with zero phase winding. 113

4.3 Plots of the polar vortex amplitudes along the x̂ axis in the polar phase

at T = 0.95Tc , p = 15bar. It is clear that the amplitude drops to zero for

the spin-up channel as expected since it hosts a singly-quantized vortex.114

4.4 The spin-polarization along the nematic axis is shown for a single HQV

vortex in the polar phase at T = 0.95Tc , p = 15bar. The magnetization

is maximum in magnitude at the core of the HQV. 115

4.5 It is clear that the asymptotic calculation |C+0 (x ⇠(T ))| ⇠


p
p 1 1/x2 is valid for x 4⇠(T ) ⇡ 400nm. 120

4.6 Amplitudes and phases of a single polar half-quantum vortex at

p = 15bar, T = 0.95Tc , on a square grid of ±800nm ⇡ 7⇠(T ). The core

size is ⇡ 30nm. The HQV is a singly-quantized vortex in the spin-up

condensate, and a constant amplitude with zero phase winding in the

spin-down condensate. The core is ferromagnetic superfluid -phase. 121

4.7 Transverse mass currents taken at p = 15bar, T = 0.95Tc . 122

P
4.8 The superfluid density, ↵,i |A↵i |2 = |Axz |2 + |Ayz |2 , is shown above

for the HQV pair at p = 15bar, T = 0.95Tc . The pair of HQVs break

rotational invariance along the x̂ axis and have superfluid, spin-polarized

ferromagnetic cores of size r ⇡ 50nm. 124


20

4.9 Amplitude of a polar HQV pair at p = 15bar, T = 0.95Tc along

the x̂ axis. It is clear that a pair of HQVs can be described as a

singly-quantized vortex in both the spin-up and down condensate as

both of these amplitudes vanish in the vortex cores. 125

4.10 Vortex magnetization profile along the ẑ axis m(r) for the HQVs at

p = 15bar, T = 0.95Tc . The spin-polarization is maximum in magnitude

in the vortex cores. Inset: Density plot of the same. 128

4.11 Diagonal components of the zero-field transverse susceptibility for the

HQVs at p = 15bar, T = 0.95Tc in units of the normal-state N. As

shown in Eq. 4.23 , we have zz = N, thus we need to only consider

the components in plane. As we move away from the vortex cores along

the x̂ axis, the vortices are more easily polarized which is expected as

superfluidity gets stronger as we move away. In the vortex cores, the

vortices are equally polarizable along either field direction and thus

xx (x, 0) ⇡ yy (x, 0). The o↵-diagonal are negligible compared with the

diagonal components and satisfy xy = yx . 129

4.12 Expanded view of the mass currents taken at p = 15bar, T = 0.95Tc

clearly showing the broken axial symmetry of the vortex. It can be

seen that the vortex structure near the half-quantum vortex cores is

the source of the anisotropic current density. The currents vanish near

the zero-phase winding polar cores. Currents are scaled in units of j0

defined by Eqn. 13. 130


21

4.13 Spontaneous zero-field magnetization as a function of the intervortex

spacing set by the initial condition of a pure polar HQV. The

magnetization is generated by the Barnett e↵ect, i.e. rotation induced.

It is clear that the Barnett e↵ect vanishes for a singly-quantized vortex.

Note, the Barnett induced magnetization of the vortices in the polar

phase are 4-5 orders of magnitude smaller than the normal-core vortex

in the Bulk B-phase as described in detail in the text. 132

4.14 The collapse of a HQV pair into a conventional singly-quantized vortex.

Both spin-up and down components drop to zero in the core and have

2⇡ phase winding as expected. 133

4.15 It is clear that as the intervortex separation x increases, the HQV

pair energy is lowered. A singly-quantized vortex, x = 0, is higher

in energy than a pair of HQVs and thus the HQV is not a metastable

phase, but a local minimum of the free energy. For x > 1000nm, the

energy starts to become larger as the stable phase is now the B phase

and no longer the confined polar phase. 134

4.16 Evolution of the polar component, <[Axz ](x, 0), along the vortex core

axis x̂ in a transverse magnetic field, Ĥ = x̂. The magnetic energy is

minimized for d̂ ? Ĥ, hence the suppression of the order parameter.

For H > 1000G, the HQV pair approaches that of a singly-quantized

vortex centered near x = 0 as shown. 136

4.17 Order parameter amplitude and spontaneous supercurrents generated

by the polar-distorted chiral phase confined in Nafen. Here, we initialize


22

a chiral phase of the form pz + ipx . The spontaneous supercurrents are

generated by the coupling of p̂x , p̂z and integrate to zero as expected. 138

4.18 Order parameter amplitude and spontaneous supercurrents generated

by the polar-distorted chiral phase confined in Nafen. Here, we initialize

a chiral phase of the form pz + i(px + py ). The spontaneous supercurrents

are generated by the coupling of p̂x , p̂y , p̂z and again integrate to zero as

expected. 139

4.19 Vortex magnetization along the ẑ axis m(r) for the HQVs at p = 15bar,

T = 0.85Tc as the transverse magnetization is zero since it couples to

d̂ = ẑ, but d̂ ? ẑ. The magnitude of the spin-polarization is maximal in

the cores. It is clear that the vortex cores are superfluid and magnetic.

Inset: Density plot of the same. 143

4.20 Intrinsic orbital angular momentum, Lk , for the HQVs at p = 15bar,

T = 0.85Tc . We note that Lk is on the order of Lorb (n~/4)( /Ef )2 ,

with orb ⇠ O(ln(Ef /kB Tc )). Inset: Density plot of Lz projecting out

the components with orbital angular momentum x̂i , ŷi indicating the

chiral nature of the vortex core Cooper pairs. 144

4.21 Anisotropic mass currents taken at p = 15bar, T = 0.85Tc . The

transverse mass currents are given by j(r) = j0 = [A⇤↵z rx A↵z x̂ + A⇤↵z ry A↵z ŷ] .145

4.22 Vortex phase diagram for the equal-spin pairing phases of confined

superfluid 3 He. We calculate the structure of a polar HQV pair in the

polar phase at p = 15bar, T = 0.95Tc and then cool in temperature into


23

the polar-distorted chiral A phase to stabilize a pair of polar-distorted

chiral HQVs at p = 15bar, T = 0.85Tc . The vortex phases are calculated

in equilibrium and compared with the experimental equilibrium phase

transitions [7]. Insets: Polar-distorted chiral HQV pair and Polar HQV

pair Cooper pairing densities. 147

4.23 Polar phase order parameter suppressed by a lattice of square impurities


p
in Nafen-90 in units of bulk P = |↵|/2 12345 at T = 0.95Tc , p = 15

bar. The impurities for Nafen-90 have interstrand distance of r ⇡ 50nm

, and the scattering cross-section eigenvalues are given by !? = 4.25nm,

!k = 1.073nm [8]. 148

4.24 Amplitudes, phases and transverse mass currents of a polar HQV pair

at p = 15bar, T = 0.95Tc , on a square grid of ±800nm ⇡ 7⇠(T ). A

pair of HQVs can be described as a singly-quantized vortex in both the

spin- up and down condensate, thus these are the only two non-zero

components of the order parameter as shown. The suppression of

the order parameter outside the vortex cores is from the Nafen-90

impurities. 150

4.25 Amplitudes, phases and transverse mass currents of a polar-distorted

chiral HQV pair at p = 15bar, T = 0.85Tc , on a square grid of

±800nm ⇡ 13⇠(T ). It is clear that this pair of HQVs nucleate a chiral

phase in its core as shown in the C++ , C + components. The dominant

bulk components are the polar components, C+0 , C 0 , as expected.

The components with d̂ k ẑ are zero since the HQVs are an equal-spin
24

pairing phase. The vortex cores are less circular than the pure polar

HQV pair and more elliptical which is a manifestation of the underlying

chiral phase order parameter on the impurity lattice. 151

4.26 Spontaneous supercurrent flowing along the polar axis ẑ, in units of j0 ,

despite having zero phase gradient along this axis. The supercurrent

integrates to zero as expected as there is no mass transport along this

axis. These currents were calculated at p = 15bar, T = 0.85T c. 153

4.27 Vortex magnetization profile along the ẑ axis m(r) for the HQVs.

The polar HQV pair is taken at p = 15bar, T = 0.95T c, while

the chiral pair of HQVs is taken at p = 15bar, T = 0.85Tc . The

spin-polarization is maximum in magnitude in the vortex cores. The

transverse magnetizations are zero since they couple to dˆ = ẑ, but

dˆ ? ẑ. It is clear that the vortex cores are superfluid and magnetic.

Inset: Density plots of the same. 155

5.1 Zero-energy local density of states peaked in the vortex cores along the

x̂ axis. Plot taken at p = 15bar, T = 0.95Tc . Inset: Cooper pairing

density for a pair of HQVs. 160

5.2 Spin-resolved zero-energy local density of states peaked in the spin-up

condensate along the x̂ axis. Plot taken at p = 15bar, T = 0.95Tc . 161

5.3 Spin-resolved zero-energy local density of states peaked in the spin-down

condensate along the x̂ axis. Plot taken at p = 15bar, T = 0.95Tc . 162


25

CHAPTER 1

Introduction

Physics is the study of the laws of nature [9]. As theoretical physicists, we are strictly

concerned with understanding the laws of the universe by using mathematical concepts to

describe nature around us. In this dissertation, I develop theoretical models to describe

quantized vortices that are indeed observed experimentally in nature. A vortex in a fluid

can be simply thought of as a whirlpool. Similarly to how we observe whirlpools in a

swimming pool or ocean, we also observe vortices in fluids that are a bit colder than the

ocean. Vortices formed at such extreme temperatures near zero Kelvin behave di↵erently

than the usual whirlpools we observe and are described in the language of quantum field

theory. In this thesis, I investigate vortices in fluids that are close to T = 0K, specifically

quantized vortices in superfluid 3 He. Surprisingly, liquid helium never solidifies at low

temperature under its own vapor pressure whereas all other liquids do [10, 11]. This

is a consequence of the interatomic forces and the large zero-point motion of the helium

atoms. You would expect that at T = 0K all motion stops, but this is only true on

a classical scale and is not correct quantum mechanically [12]. Even at T = 0K, the

helium particles refuse to settle down. Rather they stay in motion due to the quantum

mechanical zero-point energy. Since the helium atom is of very low mass, this allows

for a substantially large zero-point energy enforced by confinement and the Heisenberg

uncertainty principle. Thus Helium is the only noble gas to remain liquid at ambient

pressure down to zero temperature.


26

This exotic phenomena in Helium is known as superfluidity. Fritz London was the

first theorist to understand the connection between Bose-Einstein condensation and su-

perfluidity in Helium [13]. It was Feynman and Onsager that independently predicted

quantized circulation and vortices in liquid helium [14, 15]. It is clear that liquid he-

lium is a quite complicated quantum mechanical system [16], and it was Feynman who

realized this early on by writing five papers alone on liquid Helium in less than five years

[17, 18, 19, 15, 20].

Superfluidity is an exotic state of matter in which the fluid flows without friction.

Imagine attempting to cross a street in Times Square, you would bump into many people

as you were doing so - this is the so-called ‘normal state’ of the fluid. Now imagine

everybody crossing the street in unison lock-step together. Crossing the street would then

occur smoothly and you would experience no bumping of arms or friction. This is the

superfluid state. It is a collective state observed in nature in which all the people march as

one, or all the particles behave as one. Essentially, all of the microscopic particles form a

macroscopic single state, and this is the pure friction-less superfluid state. There are two

isotopes of helium, Helium-3 and Helium-4 which both behave as superfluids below their

specific transition temperatures. Helium-4 is a weakly interacting inert bosonic gas that

enters the superfluid state under a critical temperature of Tc ⇡ 2K. However, Helium-3

has a Fermi temperature on the order of Tf ⇠ K, and enters the superfluid state below

Tc ⇡ 2mK. Experimentally, 4 He was observed to become a superfluid in 1938 [21, 22],

while it took until 1972 to discover 3 He going superfluid [23].


27

Both superfluids, 3 He, and 4 He, host quantized vortices. However, the superfluid

phase of Helium-4 is considerably simpler and only hosts a conventional s-wave quantum

vortex described by a scalar order parameter with zero amplitude in the vortex core. This

is simply because 4 He is a s-wave, spin-singlet superfluid, (s = 0, l = 0) whereas 3 He

is p-wave spin-triplet (s = 1, l = 1) superfluid. The normal state of 3 He is a strongly-

correlated Fermi liquid with strong ferromagnetic spin fluctuations. It is these fluctuations

that directly provide the pairing “glue” that binds spin-triplet Cooper pairs. This glue

is what allows the particles to all collapse into a single state, behaving as one, and thus

allowing the superfluid state to exist. Thus, superfluidity in 3 He occurs around T ⇡ 2mK

and is the only system known to be described by spin-triplet unconventional pairing ‘su-

perconductivity’. It is miraculous that we even still have a fluid at such low temperatures!

Fluids under rotation at such low temperatures form exotic phases of matter, and here

the primary investigation was based on the formation of vortices in superfluid 3 He.

Experiments on fluids near zero Kelvin under rotation started out in the 1980s in the

rotating cryostat at Helsinki Laboratory [24, 25, 26], and has rapidly expanded globally

since. The first experiments on superfluid rotating 3 He succeeded in rotating the fluid at

a few radians per second while keeping the fluids at extremely low temperatures.

It is important to briefly discuss why and how vortices are indeed generated under

rotation in a superfluid. Typically, classical fluids co-rotate with the walls of the rotating

vessel by the viscous forces and forces of the walls on the fluid. However, superfluid 3 He is

described quantum mechanically and thus quite di↵erent. The collective motion of all the

atoms collapse into a single state described by a single wavefunction. This is analogous

to the people walking across the street in unison lock-step I described earlier. It is this
28

collective motion of the particles that leads to superfluidity. A superfluid cannot simply

rotate with the walls of its rotating container as its velocity field is irrotational. Thus,

if we rotate a vessel containing a mixture of normal and superfluid 3 He the normal fluid

will rotate with the vessel, but a superfluid will remain at rest, at least at low rotation

speeds. However, above a lower critical rotation speed, ⌦c1 , the irrotational state of the

superfluid is no longer energetically favorable. The ground state of a superfluid in a vessel

rotating with speed ⌦ > ⌦c1 is an array of vortices with quantized circulation penetrating

the fluid. Each vortex circulates in the vessel with a quantized circulation  = h/M

where h is Planck’s constant and M is the mass of the particles that condense to form

the superfluid [14, 20]. Indeed, this rotation-induced vortex formation phenomenon is

observed in both superfluid 3 He and 4 He [27, 25].

Rotation is one means of generating quantized vortices in a superfluid. However, quan-

tized vortices were recently observed to form in superfluid 3 He even under zero rotation

[28]! Half-quantum vortices were generated in the confined phases of superfluid 3 He in

Nafen without (and with) rotation [28]. The vortices generated under zero rotation are

attributed to the Kibble-Zurek mechanism [29, 30, 31]: topological defects being formed

from cooling rapidly through a second-order phase transition from the normal state into

the superfluid state.

The theoretical investigation of quantized vortices in superfluid 3 He-B in Chapter

3 explains experimental findings which indicate a phase transition between vortices in

rotating 3 He-B at temperatures below 1-2 mK. I calculated phase transition lines in the

pressure-temperature plane between two di↵erent vortex states observed in the rotating

fluid. This phase diagram was obtained from the stability of di↵erent stationary states
29

of the free energy investigated as a function of pressure, temperature and magnetic field.

This includes both equilibrium and metastable vortex phases.

In Chapter 4, I report results for the vortex phases of a theoretical model for the

superfluid phases of 3 He infused into high-porosity Nafen aerogel [8]. I report calculations

of the stability of vortex states and predict a phase transition between di↵erent phases

of superfluid 3 He confined within the Nafen aerogel that are not realized in pure 3 He

as a function of temperature, pressure and magnetic field. Excellent agreement with

experimental data on the phase diagram of 3 He in Nafen is obtained [7]. I find energetically

stable half-quantum vortices in each of the superfluid phases of 3 He confined in Nafen.

The two equilibrium phases of superfluid 3 He in Nafen host di↵erent half-quantum

quantized vortices depending on the pressure, temperature and magnetic field. Thus, as in

pure 3 He-B there are multiple equilibrium phases of rotating 3 He in Nafen. The di↵erences

between these phases is described by the symmetries and wavefunction structures of the

di↵erent quantized vortices.

These studies provide the first quantitative theory accounting for the pressure- temperature-

magnetic field phase diagram of these remarkable phases of rotating 3 He-B and 3 He con-

fined in Nafen. The tools developed in our study open the door for wide ranging studies

of superfluids, cold atomic gases to superconductors.


30

CHAPTER 2

Introduction to Quasiclassical Theory

The Bardeen-Cooper-Schrie↵er (BCS) microscopic theory of superconductivity in 1957

[32] was explained by the formation of a bound state of two fermions with opposite

momenta, one spin-up electron, and one spin-down electron. The bound-state formation

was theoretically understood by Leon Cooper in 1956 [33], and is thus called the Cooper

pair. The BCS theory then derived a ground-state wavefunction of the superconductor

based on the Cooper pairing theory. The Cooper pair condensate was derived by Cooper

by simply solving the Schrödinger equation in momentum space for two electrons with

a short-range attractive interaction between them. The unusual attraction between two

electrons was later realized to be a direct consequence of electron-phonon coupling. We

can write the Cooper-pair condensate wavefunction in terms of the number of particles,

N , and volume, V , as
r
N
(2.1) (~r1 , ~r2 ) = h " (~
r1 ) # (~r2 )i = f (|~r1 ~r2 |) ,
2V

where f (r) ⇠ O(1/⇠ 3/2 ) for r . ⇠ where ⇠ is the radial size of a Cooper pair which is

generally several orders of magnitude larger than the interatomic distance between atomic

constituents, e.g. 3 He atoms in superfluid 3 He. This separation of length scales leads to

a key simplification of the BCS formulation of the theory of superconductivity.


31

A quasiclassical formulation of the theory of superconductivity was developed by Eilen-

berger [34], and Larkin and Ovchinnikov [35] to simplfy the structure of Gorkov’s equa-

tions for BCS superconductors. The quasiclassical reduction of Gorkov’s equations are

transport-like equations defined along classical trajectories in space that are indexed by

points on the Fermi surface of the parent normal metallic state. The quasiclassical (QC)

theory is a powerful theoretical tool in condensed matter physics. The quasiclassical

equations describe the transport and evolution of 4 ⇥ 4 density matrices for the quantum

mechanical degrees of freedom - particle-hole and spin states - along classical trajectories

in phase space, with momentum confined to the Fermi surface. The central objects in

the QC theory are the Green’s function, G, and self-energy, ⌃, that encodes interactions

of fermionic quasiparticles with impurities, phonons and other quasiparticles. The quasi-

classical equations of superconductivity are derived from the full Nambu-Gorkov Green’s

function by integrating over the excitation energy which averages out the atomic scale

information of the excitations. This is valid since the Fermi wavelength, F, is usually

orders of magnitude smaller than the coherence length, ⇠0 . The QC formalism is briefly

described below.

2.1. Gorkov’s Equation

In 1958, Gorkov proposed a theory of superconductivity [36] based on the physical

idea of Cooper [33] that provided the groundwork for the calculating the properties of

a system of Fermi particles with an attractive interaction. A year later, he then showed

that this theory of superconductivity reduced to the Ginzburg-Landau equations near

the transition temperature [36]. The breaking of U (1) symmetry leads to mixing of the
32

particle and hole sectors [32, 36, 37]. Following Nambu we introduce a two-component

field that that includes both particle and hole degrees of freedom. The Nambu spinors

are then
0 1 0 1

" (~
r, t)C " (~
r, t)C
(2.2) ˆ(~r, t) = B
@ A, ˆ† (~r, t) = B
@ A.

# (~
r, t) # (~
r, t)

Now we can generalize this to a four-component spinor by writing


0 1 0 1

B " (~
r, t)C B " (~
r, t)C
0 1 B C 0 1 B C
ˆ(~r, t) B C ˆ† (~r, t) B † C
B C B # (~
r, t)C B C B # (~
r, t)C
A=B C, A=B C.

(2.3) (~r, t) = @ B C (~r, t) = @ B C
ˆ (~r, t)
† B †
r, t)C ˆ(~r, t) B r, t)C
B " (~ C B " (~ C
@ A @ A

# (~
r, t) # (~
r, t)

The quantum mechanics is directly encoded into the Hilbert space of the particle-hole
0
sector given by the s. The fermionic field operators obey anti-commutation relations

given by

n o n o
(2.4) ˆ a, ˆ † = ab ,
ˆ a , ˆ b = 0̂ .
b

Thus we can now write the 2 ⇥ 2 matrices in particle-hole space by constructing the outer

product of these four-component spinors to obtain


0 1
ˆ(1) ˆ† (2) ˆ(1) ˆ(2)
(2.5) ˆ (1) ˆ † (2) = B
@
C
A
ˆ† (1) ˆ† (2) ˆ† (1) ˆ(2)

where 1 ⌘ (~r1 , t1 , ↵1 ), 2 ⌘ (~r2 , t2 , ↵2 ) and ↵ represents spin-up and down, ", #. Now

that we have the structure of the Nambu space matrix in terms of the spinors, , we can
33

directly define a relation between the Green’s functions and Nambu spinors. The retarded

(R), advanced (A), Keldysh (K), and Matsubara (M) propagators can now be defined as

(2.6) bR (1, 2) =
G i⇥(t1 t2 )h{ (~r1 , t1 ), †
(~r2 , t2 )}i,

(2.7) bA (1, 2) = i⇥(t2


G t1 )h{ (~r1 , t1 ), †
(~r2 , t2 )}i,

(2.8) bK (1, 2) =
G ih[ (~r1 , t1 ), †
(~r2 , t2 )]i,

(2.9) bM (1, 2) =
G hT⌧ (~r1 , t1 ) ¯ (~r2 , t2 )i .

b B]
The commutator and anti-commutator are defined by [A, b = A
bBb b A,
B b {A,
b B}
b =

bB
A b +B
b A,
b respectively. The thermal expectation values are taken in the Grand Canonical
H µN H µN
ensemble and defined by h...i = Tr(e ...)/Tr(e ). The Matsubara propagator

is defined by the time-ordering operator, T⌧ , in terms of the imaginary time variable,

t 7! i⌧ for  ⌧  . The adjoint operator is then defined on the imaginary strip as


¯ (⌧ ) ⌘ †
( ⌧ ) [38]. We follow the standard notation used in describing the QC theory

for superfluid 3 He [39]. The Nambu-Gorkov matrix can then be represented as a 2 ⇥ 2

matrix in particle-hole space in terms of these Green’s functions,


0 1
x x
G (1, 2) F (1, 2)C
(2.10) bx (1, 2) = B
G @ A
x x
F̄ (1, 2) Ḡ (1, 2)

for x = R, A, K, M , where each matrix element is a 2⇥2 matrix in spin space. These prop-

agators encode both particle-hole and conjugation symmetry, and fermion anti-symmetrization.

Here, the quasiparticle diagonal Green’s functions are given by G(1, 2), Ḡ(1, 2) and the
34

anomalous Gorkov propagators are given by F (1, 2), F̄ (1, 2) describing the pairing corre-

lations of the superconducting state. We can Fourier transform this matrix in imaginary

time and transform to center of mass and relative coordinates by writing, ⌧ = it, =
~ = (~r1 + ~r2 )/2 to obtain
(kB T ) 1 , R

Z Z Z
(2.11) b p, R,
G(~ ~ ✏n ) = d⌧ e i✏n ⌧ /~
d re3 p·~
i~ r/~ ~ + ~r/2, ⌧ ; R
Ĝ(R ~ ~r/2, 0) .
0

This expression will be most useful when the momentum can be approximated by the

momentum at the Fermi surface, |~p| ⇡ pF and thus |~p| ⌧ pF . The idea from here is
b and then turn these into transport
that one can develop the equations of motion for G

(Eilenberger) equations.

2.2. Quasiclassical Green’s Function

We introduce the quasiclassical Green’s function by integrating over the Gorkov Green’s

function with respect to the normal-state excitation energy [34], ⇠p = |~vF |(p |~pF |),
0 1
Z 1 g fC
1 1 b n , p̂F , ~r) ⌘ B
(2.12) gb(✏n , p̂F , ~r) = d⇠p ⌧b3 G(✏ @ A,
a ⇡
1 f¯⇤ ḡ ⇤

g )2 =
where gb satisfies the normalization condition (b ⇡ 2b
1. The momentum dependence

in the Green’s functions arise from ⇠p , and are peaked at the Fermi surface, thus making

p~ = p̂f p̂ a reasonable valid approximation. Here, ⌧b3 is a Pauli matrix in particle-hole

space,
0 1
B1̂ 0̂ C
(2.13) ⌧b3 = @ A,
0̂ 1̂
35

where 1̂ is the unit matrix in spin space. The QC propagators are renormalized by the

spectral weight of the normal-state quasiparticle pole, 0 < a < 1. If we now write the

di↵erence of the left and right Dyson equations as

(2.14) [i✏n ⌧b3 b ⌧3


⌃b b + i~ p~F · r(b
b , ⌧b3 G] b =b
~ ⌧3 G) 0,
m

b is the self-energy matrix in Nambu space.


we can obtain the transport equation where ⌃

We describe this in more detail below.


b to be
The quasiclassical theory now approximates the momentum dependence of G

dominant near the Fermi surface, |p̂| ⇡ pF . We post multiply the self-energy by ⌧b3 , the
b ⌧3 as ⌃
redefine ⌃b b and use Eq. (2.12) to obtain the transport equation for the quasiclassical

Nambu propagator, gb, defined along classical trajectories in momentum space. This leads

to Eilenberger’s quasiclassical transport equation given by

(2.15) [i✏n ⌧b3 b (~pF , R)


~ b pF , R,
⌃(~ ~ ✏n ), gb(~pF , R,
~ ✏n )] + i~~vF · rb
~ g (~pF , R,
~ ✏n ) = 0̂ ,

where the self-energies are explicitly given by


0 1
ˆ pF , R,
⌃(~ ~ ✏n ) ˆ (~pF , R,
~ ✏n )
(2.16) ⌃(~ ~ ✏n ) = B
b pF , R, @
C
A,
ˆ ⇤ (~pF , R, ˆ¯ p , R,
~ ✏n ) ⌃(~ ~ ✏n )
F

0 1
0 ˆ mf (~pF , R)
~
(2.17) ~ =B
b (~pF , R) @
C
A.
ˆ ⇤ (~pF , R)
~ 0
mf
36

It is often useful to separately indicate the mean-field (mf) pairing self-energy as b ,

which we identify above as the order parameter for the superfluid state. This is purely
b The
an o↵-diagonal matrix. All other self energies and interactions are contained in ⌃.
b represent Fermi liquid energies, impurity and quasiparticle
diagonal components of ⌃
b describe the elastic and inelastic scattering
energies. The o↵-diagonal components of ⌃

corrections to the mean-field order parameter. The Eilenberger equation, Eq. (2.15), is
~ and the linear time-
called a transport equation because of the di↵erential operator ~vF · r

derivative operator which is represented by i✏n ⌧b3 after Fourier transformation. It can be

solved along trajectories where the flow is transported along the direction of v̂F following

the gradient r̂, analogous to the Boltzmann transport equation. Thus we can see that the

quasiclassical propagators depend on the Fermi momentum, p̂F , and position on the Fermi

surface. In the normal state, the Eilenberger equation describes the ballistic transport

of quasiparticle wavepackets propagating at the Fermi velocity. The pairing self-energy

is defined by ˆ , and the particle-hole coherence is encoded in the quasiclassical Nambu

Green’s function, gb.

The quasiclassical Green’s function for the superconducting state in non-equilibrium

is represented by a 2 ⇥ 2 matrix in Keldysh space [40],


0 1
b R ~ b K ~
G (~p, R, ✏, t) G (~p, R, ✏, t)C
(2.18) ~ ✏, t) = B
Ǧ(~p, R, @ A
0 bA (~p, R,
G ~ ✏, t)
37

bR,A,K are 4⇥4 matrices for the retarded, advanced and Keldysh Green’s functions
where G
bR,A encode spectral in-
in Nambu space. The retarded and advanced Green’s functions G

formation for quasiparticle excitations and Cooper Pair excitations. The Keldysh Green’s
bK encodes non-equilibrium phenomena for the occupation of states.
function [40], G

The non-equilibrium Eilenberger equations describe the dynamics of the QC Green’s

function, and all one-body physical observables are represented by integrals of these QC

propagators over the Fermi surface and excitation energy. This formalism was derived

in the normal state by Keldysh [40] and a di↵erent, equivalent formulation developed by

Baym and Kadano↵ [41, 42].

The propagators contain the Green’s functions in Nambu space which can be repre-

sented by
0 1 0 1
ĝ R,A ˆR,A
f ĝ K ˆK
f
(2.19) bR,A = B
G @
C
A, bK = B
G @
C
A,
fˆ˜R,A g̃ˆR,A fˆ˜K g̃ˆK

0 1 0 1
ˆ R,A ˆ R,A ˆK ˆK
B⌃ C B ⌃ C
(2.20) bR,A =@ A,
K
b =@ A.
ˆ˜ R,A ⌃
ˆ˜ R,A ˆ˜ K ˆ˜ K

The particle-hole and conjugation symmetries along with fermion anti-symmetrization re-
ˆ p, R;
late the upper and lower elements in Nambu space [43, 38] by Q̃(~ ~ ✏, t) = Q̂( p~, R;
~ ✏, t)⇤ .

ˆ while the Cooper pair propagators


The quasiclassical propagators are denoted by ĝ, g̃;

are given by the o↵-diagonal terms fˆ, fˆ˜. The particle-hole structure of the propaga-

tors (R, A, K) are spin matrices that can be further decomposed into spin-singlet scalar

components and spin-triplet vector components by writing the diagonal quasiclassical


38

propagators, g, ḡ and the anomalous propagators, f, f¯ as sums of a scalar and vector

contribution, we then obtain


0 1
g + ~g · ~ fi + f~ · (i~ y )C
(2.21) ~ ✏, t) = B
ĝ(~pf , R; @
y
A.
f˜i ~˜ ~g̃ ·
y + f · (i y ~ ) g̃ y~ y

2.3. Eilenberger Transport Equations

We can write the equilibrium transport equations in terms of just the retarded and
b Eilenberger’s
advanced Green’s function. Neglecting the self energies contained in ⌃,

transport equation, Eq. (2.15), reduces to [34]

h i
(2.22) ✏ R,A
⌧ˆ3 ˆ (~pF , R),
~ ĝ R,A (~p, R;
~ ✏) + i~~vF · rĝ
~ R,A (~p, R;
~ ✏) = 0,
h i2
(2.23) ĝ R,A ~ ✏)
(~p, R; = ⇡ 2 1̂ .

The barred functions are defined by particle-hole symmetry and satisfy ĝ¯R (~pF , R;
~ ✏) =

~ ✏)]⇤ .
[ĝ R ( p~F , R;

The o↵-diagonal, mean-field self energy is defined by the 2 ⇥ 2 matrix in spin space

Z |✏0n |<⌦c
X
(2.24) ↵ (~pF , ~r) = NF d2 p~0F pF , p~0F )kB T
↵ , ⇢ (~ f ⇢ (✏0n , p~0F , ~r)
✏0n

where the pairing interaction is given by (~pF , p~0F ), and ⌦c is the cut-o↵ energy. The ho-

mogeneous solution to the equilibrium retarded and advanced Green’s functions satisfying

Eq. (2.22),( 2.23) is given by

✏ˆ
⌧3 ˆ (~p)
(2.25) ĝ R,A (~p; ✏) = ⇡q .
(✏ ± i0)2 | ˆ (~p)|2
39

It can be shown that the equilibrium propagator satisfies the normalization condition,

[ĝ R,A ]2 = ⇡ 2 1̂ [34].

2.4. Spectral Density and Currents

These propagators encode the spectral information about the fermionic single-particle

states as well as the bosonic Cooper pair states. The single-particle local density of states

(LDOS) for the quasiparticle propagators is obtained from the 4 ⇥ 4 Green’s function in

Nambu space as

1
(2.26) N (✏, p~F , ~r) = = T r ⌧b3 gbR (✏, p~F , ~r) .
2⇡

Using the decomposition given in Eq. (2.21), it is often easiest to calculate the LDOS

using the scalar spin-singlet fermionic Green’s function

1
(2.27) N (✏, p~F , ~r) = = g R (✏, p~F , ~r) .

Similarly, the spectral density for spin-triplet Cooper-pairing correlations is given by

~ p~F , ~r) = 1 ⇣ ~R ⌘
(2.28) P(✏, = f (✏, p~F , ~r) .

The mass supercurrents generated from the orbital motion of the Cooper pairs is then

calculated directly from the density of states, or from the Green’s function as [44]

Z Z 1
(2.29) ~j(~r) = Nf 2
d p~F d✏~vf (~pF )N (✏, p~F , ~r)(2f (✏) 1) ,
1
40

which can be transformed to

Z X
(2.30) ~j(~r) = 2Nf d2 p~F ~vF (~pF )kB T g M ("n , p~F , ~r) ,
✏n

~ is the spin-scalar component of the diagonal Matsubara propagator,


where g M ("n , p~F , R)
~ by analytic continuation ✏ ! i"n where "n = (2n + 1)⇡kB T are
related to g R ("n , p~F , R)

the Fermion Matsubara energies. The latter representation is more efficient for computing

equilibrium properties.

There are also spin currents generated within the vortices arising from the intrinsic

spin of the Cooper pairs. The spin-current spectral function can be written in terms of

the quasiclassical propagator as

~ p~F , ~r) = 1
(2.31) S(✏, = ~g R (✏, p~F , ~r) .

Thus, the intrinsic spin-current LDOS can then be obtained as

(2.32) ~j↵ (✏, p~F , ~r) = 2Nf ~ ~vp (S↵ (✏, p~F , ~r) S↵ (✏, p~F , ~r))
2

where ~j↵ is the ↵ component of the spin-current spectral density flowing along the v̂p

direction [45].

When computing numerical solutions to the Eilenberger equation, it is typically more

convenient to work in the Matsubara representation [46] in which the Green’s func-

tions are calculated for energies at discrete values along the imaginary axis given by

GR (~p, R; bM (~p, R;
~ ✏) = limi✏ !✏+i0+ G bA (~p, R;
~ ✏n ), G ~ ✏) = limi✏ !✏ i0+
bM (~p, R;
G ~ ✏n ), which re-
n n

lates the retarded (advanced) parts to the Matsubara part by analytical continuation from
41

the upper (lower) half of the complex energy plane to the real axis. Note that the Green’s

functions in momentum and Matsubara frequency space are related to the real- space and

time Green’s functions by a Fourier transform

Z Z Z
(2.33) b p, R;
G(~ ~ ✏n ) = d⌧ e i✏n ⌧ /~ 3
d re p·~
i~ r/~ ~ + ~r/2, ⌧ ; R
Ĝ(R ~ ~
R/2, 0)
0

~ = (~r1 + ~r2 )/2, ~r = ~r1


where R ~r2 are the center of mass and relative coordinates, respec-

tively. The quasiclassical approach is a powerful technique to compute both equilibrium

and non-equilibrium dynamical properties of superfluids and superconductors. This is at-

tributed to the fact that the QC theory is well suitable for numerical calculations allowing

one to understand many properties of superconductivity for a large variety of geometries

for both clean and dirty superconductors.

2.5. Riccati Transformation

It is often more convenient to solve the Eilenberger equation numerically by first

transforming it into a Riccati non-linear di↵erential equation [47]. This transformation

avoids having to deal with numerical instabilities implicit in the Eilenberger equation.

However, one can obtain analytical solutions to the Eilenberger equation directly [48].

Riccati di↵erential equations are numerically stable and obtained by a transformation

of the Green’s functions in spin space. The transport equations are then solved along

Riccati trajectories in momentum space. For a thorough derivation of the transformation

between the Eilenberger and Riccati equations, see the review by Shelankov [49]. I follow

the notation and formulation of Ref. [50] starting with the introduction of the projection
42

operators
! !
1 1̂ 1 b
1
(2.34) Pb+ = b
1+ gb , P̂ = b
1 gb
2 i⇡ 2 i⇡

which obey the algebra defining projection operators into particle (+) and hole (-) sectors,

(2.35) Pb±2 = Pb± , Pb+ Pb = Pb Pb+ = 0 .

These identities follow directly from the normalization condition, Eq. 2.23 The parame-

terization of the Green’s function can be written as


0 1
1̂ + ˆ ¯ ˆ 2ˆ C
(2.36) gb = i⇡ RbB@ A
2 ˆ¯ 1̂ ¯ˆ ˆ

where the prefactor is defined as


0 1
B(1̂ ˆ ˆ¯ ) 1
0 C
(2.37) R=@ A.
0 (1̂ ˆ¯ ˆ ) 1

Using this transformation with Eq. (2.15), we obtain the Riccati equations in terms of

Riccati amplitudes, , ¯ :

(2.38) ~ =
i~vF · rˆ 2i✏n ˆ + ˆ ˆ¯ ˆ + 2⌃ˆ
ˆ ˆ,

(2.39) ~ ˆ¯ =
i~vF · r 2i✏n ˆ¯ + ˆ¯ ˆ¯ ˆ¯ ˆ
2⌃ˆ ˆ.

The Riccati amplitudes obey particle-hole symmetry relations given by

(2.40) ~ ✏n ) = ˆ ( p̂F , R,
ˆ¯ (p̂F , R, ~ ✏n )⇤ ,
43

and are related to the quasiparticle and Cooper-pair propagators by the relations

(2.41) ˆ= (i⇡ 1̂ ĝ) 1 fˆ, ˆ¯ = (i⇡ 1̂ + ḡˆ) 1 fˆ¯ .

The connection with the projection operators in Eq. (2.34) is now obvious. The Riccati

amplitudes must be calculated along classical trajectories in momentum space, and the

solutions are then interpolated back to the real-space grid in which the Green’s function

and physical observables are then calculated on. The trajectories are one-dimensional in
~ and thus by
momentum space with direction specified through the gradient term ~vF · r,

the Fermi momentum p̂F . The initial conditions for the Riccati amplitudes are chosen

such that the numerical integration is stable, and thus in the direction of ±v̂F for ˆ , ˆ¯

respectively. The homogeneous Riccati amplitude in equilibrium is given by

(2.42) = p ,
✏2n + ✏2n + | |2

and will be the basis for the initial value for the numerical integration procedures outlined

later in Chapter 6.

2.6. P-Wave, Spin-Triplet Superfluid 3 He

For 3 He the maximal symmetry group of the normal phase in zero magnetic field is

given by

(2.43) G = SO(3)L ⇥ SO(3)S ⇥ U (1)N ⇥ P ⇥ T .


44

Here we have assumed that we can neglect the spin-orbit coupling due to the nuclear

dipole-dipole interaction, this is reasonable since the dipole interactions are not too rele-
4
vant when considering the formation of the condensed superfluid phase as they are 10

smaller than the pairing gap energy. Dipole-dipole interactions are important in resolving

degeneracies with respect to relative spin-orbit rotations. As such the are essential for

understanding the NMR spectra of the superfluid phases of 3 He [51, 52], and can be

treated perturbatively.

The superfluid phases are condensates of p-wave, spin-triplet (l = 1, s = 1) Cooper

pairs described by the pairing-gap order parameter matrix as


0 1
dx + idy dz C
(2.44) ˆ (p̂, ~r) = i~ y
~ ~r) = B
· d(p̂, @ A.
dz dx + idy

~ it defines
Here, the direction of zero-spin projection is given by <(d↵ ). For complex d,

spin-polarized Cooper pairs and can be parameterized in terms of a spin-triplet p-wave

order parameter as

(2.45) d↵ = A↵i pi .

The p-wave pairing order parameter A↵i is a 3 ⇥ 3 complex matrix where ↵, i refer to the

spin- and orbital degrees of freedom of the Cooper pairs. The orbital basis states of the

Cooper pairs are given by pi for i = {x, y, z}, i.e. the direction cosines of the relative

momentum of the pairs, p~, along the Cartesian axes, {x̂, ŷ, ẑ}.

Below, we will consider quantized vortices in the bulk B phase of superfluid 3 He and the

confined superfluid phases of 3 He in Nafen. The B Phase has an isotropic energy gap and is
45

invariant under joint spin and orbital rotations (J = 0) and thus has d~ = p̂ [53]. The A

phase of superfluid 3 He is an equal-spin pairing phase and chiral since it is invariant under

reflection symmetries and breaks time reversal symmetry. The direction of the zero-spin

projection is along ẑ and thus the gap order parameter is thus given by A↵i = ẑ↵ (x̂ + iŷ)i

and has point nodes in the gap structure unlike the isotropic fully-gapped B phase. The

A phase was stabilized only in strong-coupling [54, 55]. Perhaps, the strong-coupling

stabilization is due to strong anti-ferromagnetic spin fluctuations [56, 57]. The confined

phases of 3 He in Nafen we consider are ESP states with d~ ? ẑ. The high-temperature

phase is the polar phase with orbital state aligned along the nemtic axis ẑ, i.e. pz , while

the low-temperature phase is the polar distorted chiral phase in which the transverse

oribtal amplitudes are ±⇡/2 out of phase with the polar amplitude. The decomposition
ˆ vector for this class of ESP states into A↵i is the basis for the Ginzburg-Landau
of the d(p̂)

theory which we now describe below.

2.7. Ginzburg-Landau Expansion

In 1959, Gorkov showed that the Bardeen-Cooper-Schrie↵er theory reduces to the

Ginzburg-Landau (GL) theory near the transition temperature, T ⇡ Tc [36]. Ginzburg-

Landau theory is a powerful theory within condensed matter physics that has been widely

used to study phase transitions and superconductivity for over sixty years [58]. Ginzburg

and Landau constructed a theory that describes a superconducting phase transition in

terms of a complex wavefunction, or more commonly called the ‘order parameter’. The
46

theory is based on constructing a free-energy functional as a function of the order param-

eter, which is derived from the underlying symmetry group of the superfluid or super-

conductor. The order parameter is a complex number that describes the phase transition

from the normal phase into the superconducting phase in the vicinity of the critical tran-

sition temperature, T ⇡ Tc . We now construct the leading-order free energy functional of

superfluid 3 He in terms of the complex order parameter, A↵i . Since 3 He has spin-triplet

p-wave Cooper pairs, there are 2(2l + 1)(2s + 1) = 18 degrees of freedom. Thus the order

parameter, A↵i is represented by a 3 ⇥ 3 complex matrix. We obtain the GL free-energy

functional by constructing the leading-order invariants under the symmetry group G. The

full symmetry group of the normal phase 3 He, G, is used to construct the leading order

invariants in the Ginzburg-Landau free energy functional. Thus, we have to carefully

analyze the group symmetry given in Eq. (2.43) as

(2.46) G = SO(3)S ⇥ SO(3)L ⇥ U (1) ⇥ P ⇥ T .

Here, the discrete symmetries are represented by P, T , parity and time-reversal symme-

try. The continuous symmetries are given by SO(3)S , SO(3)L , U (1). We follow a similar

derivation as done previously [59, 60]. We first analyze the special orthogonal group in

three dimensions which is invariant under spin rotations, SO(3)S . The order parameter,

A↵i , transforms as a vector under spin rotations with respect to the spin ↵,

S2SO(3)S
(2.47) A↵i ! S↵ A i , S↵ S = ↵ .

Now we consider SO(3)L , the special orthogonal group in three dimensions invariant under

orbital rotations. The order parameter also transforms as a vector under orbital rotations
47

with respect to the orbital direction i, and can be written as

O2SO(3)L
(2.48) A↵i ! A0↵i = Oij A↵j , Oij Okj = ik .

We have shown how the order parameter transforms under spin and orbital rotations.

The order parameter is also invariant under the unitary group of degree one, U (1), and

thus we must have equal numbers of A↵i , A⇤ j . We also know that A↵i transforms as an

irreducible representation of the symmetry group G, which indicates that there is only

one second-order invariant given by

(2.49) A↵i A⇤↵i = T r(AA† ) .

The invariants are calculated by considering the continuous symmetries, SO(3)S ⇥SO(3)L ⇥

U (1), and contracting spin with spin indices, orbital indices with orbital indices, and de-

manding U (1) by an equal number of A, A⇤ . We do not mix contractions between spin-

and orbital degrees of freedom. Thus, up to fourth-order, we obtain five unique invariants

given by

(2.50) 1 : (A↵i A↵i )(A j A j )⇤ = |T r(AA> )|2

(2.51) 2 : (A↵i A⇤↵i )(A j A⇤ j ) = (T r(AA† ))2

(2.52) 3 : A↵i A i A⇤ j A⇤↵j = T r(AA> (AA> )⇤ )

(2.53) 4 : A⇤↵i A i A⇤ j A↵j = T r(AA† AA† )

(2.54) 5 : A⇤↵i A i A j A⇤↵j = T r(AA† (AA† )⇤ ) .


48

Thus, we have constructed the single second-order invariant under the group G, and the

five fourth-order invariants under the group G. We can now write the Ginzburg-Landau

free energy functional for superfluid 3 He to fourth-order to obtain

Z
Fbulk = ↵(T )T r(AA† ) + 1 |T r(AA
>
)|2 + 2 (T r(AA

))2 +
R3
>
(2.55) 3 T r(AA (AA> )⇤ ) + 4 T r(AA

AA† ) + 5 T r(AA

(AA† )⇤ ) .

This is the homogeneous bulk free-energy functional that describes the bulk superfluid

A and B phases. However, there are also kinetic energies associated with the spatial

inhomogeneous variations of the order parameter. There are many inhomogeneous phases

of matter such as quantized vortices [61], confined phases in Nafen, solitons [62, 63],

domain walls, and many others. We construct here the leading-order invariants of the

gradient terms in the long-wavelength limit by contracting spin indices with themselves,

and similarly for the orbital indices. These gradient terms are directly responsible for the

inhomogeneous phases of 3 He. We obtain three unique invariant gradient terms:

(2.56) 1 : ( irj A↵i )(irj A⇤↵i )

(2.57) 2 : ( iri A↵i )(irj A⇤↵j )

(2.58) 3 : ( irj A↵i )(iri A⇤↵j ) .

We can now write the gradient energy to leading order in terms of these three invariants:

Z
(2.59) Fgrad = 1 rj A↵i rj A⇤↵i + 2 ri A↵i rj A⇤↵j + 3 rj A↵i ri A⇤↵j .
R3
49

The full free-energy functional of superfluid 3 He is given by F = Fbulk + Fgrad . We now

take the functional derivative of the bulk and gradient terms in order to minimize the
F
functional to find the energy minima; g ⌘ A†↵i
= 0 . The functional gradient can be

written as

(2.60)

gBulk = ↵(T )A + 2 1 A⇤ T r(AA> ) + 2 2 AT r(AA† ) + 2 3 AA> A⇤ + 2 4 AA† A + 2 5 A⇤ A> A .

Now taking the functional derivative of the gradient terms and using integration by parts

we have

(2.61) ggrad = 1 rj rj A↵i (2 + 3 )ri rj A↵j .

Combining Eqs. (2.60), (2.61) we obtain a complicated non-linear coupled partial di↵er-

ential equation for the nine complex matrix elements of A↵i given by

↵(T )A + 2 1 A⇤ T r(AA> ) + 2 2 AT r(AA† ) + 2 3 AA> A⇤ + 2 4 AA† A + 2 5 A⇤ A> A

(2.62) 1 rj rj A↵i (2 + 3 )ri rj A↵j = 0 .

This equation represents 18 coupled di↵erential equations that describes the entire order

parameter space. Note that in weak coupling theory, we have the coefficients of the
50

invariants in terms of the single-spin density of states, N0 , given by [64]

✓ ◆
1 T
↵(T ) = N (0) 1
3 Tc
wc 7⇣(3) N (0)
=
240 (⇡kB TC )2
wc wc wc wc wc
2 1 = 2 = 3 = 4 = 5

1
wc wc wc
1 = 2 = 3 = N (0)⇠GL
2
5
r
7⇣(3) hvf
⇠GL = .
12 2⇡kB Tc

Plugging in for the i terms in units of wc ⌘ and since ↵ < 0, ↵(T ) ⌘ |↵(T )|, we

have

0 =  r2 A + 2ri rj A↵j + |↵|A + 2 A⇤ T r(AA> )

(2.63) 2 · 2 AT r(AA† ) 2 · 2 AA> A⇤ 2 · 2 AA† A + 2 · 2 A⇤ A> A

Since A20 = |↵|/(2 · 5 wc ), using A ⌘ A0 A, x ⌘ x/⇠ where ⇠ 2 = /|↵|, we obtain a

dimensionless Ginzburg-Landau equation as

1
0 = r2 A↵i + 2ri rj A↵j + A↵i + A⇤↵i T r(AA> )
5
2 2 2 2
(2.64) A↵i T r(AA† ) (AA> A⇤ )↵i (AA† A)↵i + (A⇤ A> A)↵i
5 5 5 5

Note this is valid for the bulk B phase since we are in units of the bulk, A20 = |↵|/(10 )

where 3 12 + 345 =5 .

Assuming translational invariance along the ẑ axis, @z = 0, we can simplify these equations
51

a bit further. We explicitly write out the equations for i = x, y, z to obtain

1
0 = r2 A↵x + 2@x (@x A↵x + @y A↵y ) + A↵x + A⇤↵x T r(AA> )
5
2 2 2 2
(2.65) A↵x T r(AA† ) (AA> A⇤ )↵x (AA† A)↵x + (A⇤ A> A)↵x
5 5 5 5
1
0 = r2 A↵y + 2@y (@x A↵x + @y A↵y ) + A↵y + A⇤↵y T r(AA> )
5
2 2 2 2
(2.66) A↵y T r(AA† ) (AA> A⇤ )↵y (AA† A)↵y + (A⇤ A> A)↵y
5 5 5 5
1 2 2
0 = r2 A↵z + A↵z + A⇤↵z T r(AA> ) A↵z T r(AA† ) (AA> A⇤ )↵z
5 5 5
2 2
(2.67) (AA† A)↵z + (A⇤ A> A)↵z
5 5

where r2 = @x2 + @y2 . Note that these equations are non-linear coupled elliptic partial

di↵erential equations for a 3 ⇥ 3 complex matrix A↵i resulting in 18 coupled di↵erential

equations. We solve this equation numerically and also analytically in later chapters. We

generalize this functional later to include magnetic fields and impurity e↵ects.
52

CHAPTER 3

The Vortex Phase Diagram of Rotating Superfluid 3 He-B

We present the first theoretical calculation of the pressure-temperature-field phase

diagram for the vortex phases of rotating superfluid 3 He-B.1 Based on a strong-coupling

Ginzburg-Landau functional that accounts for the relative stability of the bulk A and B

phases of 3 He at all pressures, we report calculations for the internal structure and free

energies of distinct broken-symmetry vortices in rotating superfluid 3 He-B. Theoretical

results for the equilibrium vortex phase diagram in zero field and an external field of

H = 284 G parallel to the rotation axis, H k ⌦ , are reported, as well as the supercooling

transition line, TV⇤ (p, H). In zero field the vortex phases of 3 He-B are separated by a first-

order phase transition line TV (p) that terminates on the bulk critical line Tc (p) at a triple

point. The low-pressure, low-temperature phase is characterized by an array of singly-

quantized vortices that spontaneously breaks axial rotation symmetry, exhibits anisotropic

vortex currents and an axial current anomaly (D-core phase). The high-pressure, high-

temperature phase is characterized by vortices with both bulk A phase and phase in

their cores (A-core phase). We show that this phase is metastable and supercools down

to a minimum temperature, TV⇤ (p, H), below which it is globally unstable to an array of

D-core vortices. For H & 60 G external magnetic fields aligned along the axis of rotation

increase the region of stability of the A-core phase of rotating 3 He-B, opening a window

1This chapter is a re-formatted version of our published paper in Physical Review B [65].
53

of stability down to low pressures. These results are compared with the experimentally

reported phase transitions in rotating 3 He-B.

3.1. Introduction

The velocity field of a superfluid is irrotational. Nevertheless superfluids can approxi-

mate solid body rotation when confined in a container rotating at constant angular speed.

Co-rotation is achieved by the nucleation of an array of vortices, each of which possesses

a quantum of circulation. In superfluid 4 He, or in a spinless, s-wave BCS superfluid the

condensate wavefunction, or order parameter, is a complex scalar field. The quantum

of circulation is then  = h/M , where h is Planck’s constant and M is the mass of the

fundamental constituent of the condensate [14, 15].

Quantization of circulation reflects the single-valuedness of the condensate wave func-

tion, and non-trivial topology of the degeneracy space of the order parameter manifold. In

a cylindrical container vortices align parallel to the angular velocity, ⌦ , and co-rotation

is achieved at an average areal vortex density of nV = 2⌦/ [15]. Long-range, repul-

sive interactions lead to a two-dimensional lattice of rectilinear vortices, which for axially

symmetric vortices is a two-dimensional hexagonal lattice with inter-vortex spacing, d,


p
determined by d2 = / 3⌦, which depends only on fundamental constants and the speed

of rotation. Thus, for 4 He, or an isotropic BCS superfluid, once a sufficient number of

axially symmetric vortices nucleate to form the vortex lattice no further symmetry break-

ing phase transition is expected until the density approaches a critical density at which

neighboring vortex cores overlap and superfluidity is destroyed at an upper critical rota-

tion speed of ⌦c2 ⇡ /⇠ 2 . For superfluid 3 He which is a BCS condensate of Cooper pairs
54

with  = h/2m3 ⇡ 0.066 mm2 /s [66] and a core size ⇠ ⇡ 20 80 nm over the pressure

range p = 0 34 bar, ⌦c2 & 107 s 1 , which is experimentally inaccessible.

However, the ground state of superfluid 3 He is a time-reversal invariant, spin-triplet,

p-wave topological superfluid that breaks orbital and spin rotation symmetries, SO(3)L ⇥

SO(3)S , in addition to U(1)N gauge symmetry, but is invariant under joint spin and or-

bital rotations, SO(3)L+S [67]. The resulting degeneracy space allows for a number of

unique topologically stable defects [68, 69], including quantized vortices with di↵erent

internal core structures [70, 71]. This opens the possibility of multiple superfluid phases

characterized by distinct vortex structures.

Indeed experimental evidence of multiple vortex phases in rotating 3 He-B was reported

soon after the first rotating milli-Kelvin cryostat in Helsinki was operational [24, 25].

Using nuclear magnetic resonance (NMR) spectroscopy the vortex array in rotating su-

perfluid 3 He-B was detected as a change in the level spacing of the spin-wave bound-

state spectrum proportional to the vortex density, !sw / nV / ⌦, for rotation speeds,

⌦ = 0.2 1.7 rad/s [25, 72]. A discontinuity in !sw /⌦ at TV⇤ ⇡ 0.6 Tc was the sig-

nature of a first-order phase transition associated with the vortex array [24, 25]. The

rotation-induced NMR bound-state frequency shift also depends on the relative orien-

tation of the NMR field and the angular velocity, i.e. there is a gyromagnetic splitting,

!gyro / nV H·MV , indicative of an intrinsic magnetization generated by the circulation of


ˆ , the magnitude
the spin-triplet Cooper pairs in the region of the vortex-core, MV = MV ⌦

of which depends on the internal structure of the vortex core [25].

There are two equilibrium phases of 3 He-B under rotation. Over most of the p-T

phase diagram rotating 3 He-B is believed to be defined by an array of line defects that are
55

singly quantized mass vortices, each of which spontaneously breaks rotational symmetry,

manifest by an anisotropic, double-core structure (D-core) of the Cooper pair density.

This structure for the low-temperature, lower pressure vortex phase was discovered by

Thuneberg based on numerical solutions of the GL equations that did not constrain the

order parameter to be axially symmetric [71]. At higher temperatures and pressures the

phase of rotating 3 He-B is believed to be an array of vortices in which local rotational

symmetry is restored, but time-reversal symmetry is broken via the nucleation of both

the chiral A phase and the non-unitary phase in the core. The stability of 3 He-B with

an array of A-core vortices with ferromagnetic cores was argued based on a symmetry

classification of axially symmetric B phase vortices and the observation of a measurable

gyromagnetic e↵ect from vortices in rotating 3 He-B by Salomaa and Volovik [70]. How-

ever, a quantitative theory of the relative stability of the A-core and D-core vortex phases

as a function of pressure, temperature and magnetic field was beyond the scope of existing

theory of superfluid 3 He until now.

Based on a recent formulation of the strong-coupling Ginzburg-Landau theory that ac-

counts for the relative stability of the bulk A and B phases of 3 He for all pressures [73, 1],

we report calculations of the internal structure and energetics of topologically distinct vor-

tices in rotating superfluid 3 He-B. In particular, we report the first theoretical calculation

of the pressure-temperature-field phase diagram for the vortex phases of rotating super-

fluid 3 He-B. Theoretical results for the equilibrium vortex phase diagram in zero-field and

in an external field of H = 284 G parallel to the rotation axis, H k ⌦, are reported, as

well as the supercooling transition, TV⇤ (p, H), defining the region of metastability of the

A-core vortex phase. Central results reported here include the equilibrium phase diagram
56

A-Core A
Equilibrium
30

25 A-Core
D-Core Metastable
Equilibrium

20 PCP
p (bar)

p vc N
15
B TAB (p) - Expt., H = 0 G

10 TAB (p) - Theory, H = 0 G


TAB (p, H) - Theory, H = 284G
TV (p, H) - Expt., On Warming, H = 284G
TV (p, H) - Expt., On Warming, H = 284G

5 TV (p, H) - Expt., On Cooling, H = 284G


TV (p) - Theory, Equilibrium, H = 0G
TV (p, H) - Theory, Equilibrium, H = 284G
TV (p, H) - Theory, Metastability, H = 284G

0
1.2 1.4 1.6 1.8 2.0 2.2 2.4
T (mK)

Figure 3.1. The vortex core transition line, TV (p, H), for H = 0 G (solid
green) separating the A-core and D-core vortex phases of 3 He-B terminates
at a triple point (pvc , Tvc ) = (18.40 bar, 2.19 mK). In a magnetic field the
vortex core transition line extends to low temperatures down to p = 0
bar in a magnetic field, as shown for H = 284 G (dashed green). For
comparison the Bulk AB transition lines for are shown in blue for H = 0 G
(solid) and H = 284 G (dashed). The A-core phase supercools down to the
metastability limit, TV⇤ (p, H), shown as the purple dashed line for H = 284
G. Experimental data for the transition on cooling (red diamonds) agrees
well with the supercooling transition, while the data point at p = 29.3
bar taken on warming (red square/circle) agrees well with the calculated
equilibrium vortex phase transition. The experimental data is from Ref. [5].
57

based on precise numerical solutions of the strong-coupling theory for the vortex phases

of rotating 3 He-B shown in Fig. 3.1, as well as the region of a metastable A-core phase.

Also shown in Fig. 3.1 are the experimental results for the first-order phase transitions

between distinct vortex phases in rotating 3 He-B, both on cooling and on warming. The

transitions on cooling for H = 284 G over a wide pressure range agree with the theoreti-

cally determined metastability transition, TV⇤ (p, H), at which the A-core phase is globally

unstable for pressures p & 20 bar. Furthermore, the transition on warming at p = 29.3 bar

and H = 284 G is in close agreement with our determination of the equilibrium transition

line, TV (p, H), at that pressure and field. We discuss the phase diagram in more detail

in Sec. 3.5. These results provide strong theoretical support for the identification of the

vortex phases of 3 He-B as those originally proposed: the low-temperature, low-pressure

D-core vortex phase by Thuneberg [71], and the high-pressure, high-temperature phase

as an array of A-core vortices. The A-core vortex phase first described by Salomaa and

Volovik was originally proposed as the low-temperature vortex phase [70].

In Sec. 3.2 we begin with a description of the strong-coupling GL theory that is the

basis for our analysis summarized in Fig. 3.1. In Sec. 3.3 we describe the stationary state

vortex solutions of the strong-coupling GL theory, including their topology and broken

symmetries. We describe the key features of the axi-symmetric A-core vortex phase as well

as the non-axi-symmetric D-core vortex, including their internal topology, mass currents

and magnetic properties. Visualization of the amplitude and phase structure of vortex

states leads us to identify the mechanism responsible for the phase transition to the D-

core phase at TV (p, H). In Sec. 3.4 we discuss the local magnetic susceptibilities of the
58

A-and D-core vortices, and the resulting field evolution of the equilibrium A-core to D-

core transition. We discuss the metastability of the A-core phase in Sec. 3.5, and the

analysis underlying the supercooling transition line, TV⇤ (p, H), shown in Fig. 3.1. Our

numerical results for the stationary states of the free energy functional are based on a fast

converging algorithm described in Appendix 3.9.

3.2. Ginzburg-Landau Theory

The B phase of superfluid 3 He is the p-wave, spin-triplet Balian-Werthamer state that

is invariant under joint spin and orbital rotations as well as time reversal, H = SO(3)L+S ⇥T.

The corresponding degeneracy space of 3 He-B allows for a unique spectrum of topologically

stable defects, including several quantized mass current vortices with distinct broken

symmetries [74, 70, 71]. Topological defects often host distinct inhomogeneous phases,

confined within their cores, but embedded in the order parameter field of the ground

state [75]. Thus, a theoretical description of vortices in rotating 3 He-B requires a theory

allowing for all possible realizations of the order parameter for spin-triplet, p-wave pairing.

The GL theory is formulated as a functional of the order parameter, the amplitude

for the condensate of Cooper pairs, h (p) 0 ( p)i in the spin-momentum basis. For

spin-triplet, p-wave Cooper pairs the condensate amplitude can be expressed in terms

of a 3 ⇥ 3 matrix order parameter, A↵i , of complex amplitudes that transforms as the

vector representation of SO(3)S with respect to the spin index ↵ = {x0 , y 0 , z 0 }, and as the

vector representation of SO(3)L with respect to the orbital momentum index i = {x, y, z}.

The GL free energy functional is expressed in terms of linearly independent invariants

constructed from A↵i , A⇤↵i and their gradients, rj A↵i and rj A⇤↵i . In particular, the GL
59

functional can be expressed in terms of free energy densities [64, 73],

Z
(3.1) F[A] = d3 r (fbulk [A] + ffield [A] + fgrad [A]) ,
V

where the bulk free energy density is given by one second-order invariant and five fourth-

order invariants,

2
fbulk [A] = ↵(T )Tr AA† + 1 Tr AAT
⇥ ⇤2
+ 2 Tr AA† + 3 Tr AAT (AAT )⇤

(3.2) + 4 Tr (AA† )2 + 5 Tr AA† (AA† )⇤ ,

where A† (AT ) is the adjoint (transpose) of A.

The nuclear Zeeman energy for spin-triplet pairs also plays a role in the determination

of the vortex structure and phase diagram for the vortex phases of rotating 3 He-B, even

for relatively weak fields. The dominant field-dependent term in the GL functional is a

bulk term representing a correction to the nuclear Zeeman energy from the condensate of

spin-triplet Cooper pairs,

(3.3) ffield [A] = gz H↵ AA† ↵


H .

Note that microscopic pairing theory implies gz > 0 [64], in which case there is a cost in

Zeeman energy for S = 1, Ms = 0 triplet pairs projected along H.


60

Spatial variations of the order parameter also incur a cost in kinetic and bending

energies described by the gradient terms,

(3.4) fgrad [A]=K1 A⇤↵j,k A↵j,k+K2 A⇤↵j,j A↵k,k+K3 A⇤↵j,k A↵k,j ,

where A↵i,j ⌘ rj A↵i . The gradient energies and related currents are discussed in more

detail in Sec. 3.3.4.

The nuclear magnetic dipole-dipole interaction energy per atom is of order, n ( ~)2 ⇠
4
10 mK. This is a very weak perturbation compared to the binding pairing energy of

Cooper pairs of order, Tc ⇠ 1 mK. Nevertheless, the dipolar energy plays a central role in

the NMR spectroscopy of the superfluid phases of 3 He, and specifically the spectroscopy

of the vortex phases of rotating 3 He-B, because the dipole energy couples the spin and

orbital degrees of freedom of the spin-triplet, p-wave condensate. Thus, in addition to the

primary contributions to the GL functional (Eqs. 3.2-3.4), the mean nuclear dipole-dipole

interaction energy, contributes to the GL functional a term second-order in the order

parameter,

⇥ ⇤
(3.5) fdipole = gD |Tr A |2 + Tr AA⇤ 2
3
Tr AA† ,

where the material parameter, gD , is determined by measurement of the slope of the

square of the longitudinal resonance frequency [64] 2, !B , for bulk 3 He-B

gD = 3
5 B
(1 + F0a ) 1
Tc (d(~!B )2 /dT |Tc ),

2For a detailed analysis of the determination of gD see Ref. [76].


61

where F0a is the exchange interaction for normal-state quasiparticles in units of the Fermi

energy per atom of 3 He, and B determines the bulk order parameter of 3 He-B (c.f.

Sec. 3.2.1). The nuclear dipole energy is too weak to a↵ect the relative stability of the

vortex phases. But, when treated perturbatively, describes the dipolar energy of textures

in rotating 3 He-B that are modified by the vortex currents and the intrinsic magnetization

generated by rotation. These hydrodynamic and hydromagnetic e↵ects are discussed in

detail in Refs. [25, 64]. Here we are interested in the internal structure and stability of

the vortices in rotating 3 He-B, and thus we can neglect the nuclear dipole energy in our

analysis of the energetics of the vortex phases.

3.2.1. Material Parameters

The material parameters, ↵, { i | i = 1 . . . 5}, gz , and {Ka | a = 1, 2, 3} multiplying the

invariants defining the GL functional, which in general are functions of temperature and

pressure, are determined by the microscopic pairing theory for 3 He [39]. The coefficient

of the second-order invariant determines the zero-field superfluid transition [64, 73],3,

(3.6) ↵(T ) = 13 Nf (T /Tc 1) ,

where Nf = m⇤ kf /2⇡ 2 ~2 is the single-spin normal-state density of states at the Fermi

level expressed in terms of the quasiparticle e↵ective mass, m⇤ , and Fermi wavenumber,

kf . The latter is determined by the particle density n = kf3 /3⇡ 2 . In addition, the Fermi

momentum, pf = ~kf , Fermi velocity, vf = pf /m⇤ and Fermi energy, Ef = 1


v p ,
2 f f

3There is a very small correction to @↵/@T |Tc from the finite lifetime of quasiparticles which has no role
in the relative stability of the vortex phases.
62

determine the GL material parameters, all of which depend on pressure via the equilibrium

particle density (see Table 3.2 of Appendix 3.8).

For the homogeneous bulk phase it is convenient to represent the order paramter

matrix in terms of an amplitude and normalized matrix, A = a where Tr aa† = 1.

Then for any stationary solution of the bulk free energy functional, Eq. 3.1, the pair
2
density is = |↵(p, T )|/2 a , where a is a local minimum of the functional [a] = 2 +

1 |Tr aaT |2 + 3 Tr aaT (aaT )⇤ + 4 Tr (aa† )2 + 5 Tr aa† (aa† )⇤ . The corresponding

bulk free energy density for the stationary solution is then, fa = 12 ↵ 2


a = 1 2
4
↵ / a.

In weak-coupling BCS theory the relative values of the five fourth-order materials

parameters are uniquely determined,

wc wc wc wc wc
(3.7) 2 1 = 2 = 3 = 4 = 5 ,

wc 7Nf ⇣(3)
(3.8) where 1 = .
240(⇡kB Tc )2

As a result the weak-coupling BCS formulation of GL theory predicts a unique bulk


p
phase, the Balian-Werthamer (BW) state [53] defined by AB ↵i = B ↵i / 3 where B =
p
|↵(T )|/2 B , which is the ground state at all pressures in zero magnetic field. The

magnitude of the B-phase order parameter is defined by B ⌘ 12 + 13 345 where ijk... =

i + j + k + . . .. In the weak-coupling theory wc


B = 53 | wc
1 |. For comparison, the bulk A

phase, first discussed as a possible ground state of 3 He by Anderson and Morel (AM) [54],
p p
is defined by AA ↵i = A ẑ↵ (x̂i + iŷi ) / 2, where A = |↵(T )|/2 A with A ⌘ 245 ,
63

wc wc
which in the weak-coupling limit becomes A = 2| 1 |. Thus, in weak-coupling theory

the A phase is never stable relative to the B phase.

For inhomogeneous states the coefficients of the gradient energies determine the re-

sponse of the order parameter to strong perturbations, e.g. the spatial variations, both

suppression and growth, of order parameter components in the cores of vortices and topo-

logical defects. In the weak-coupling limit the sti↵ness coefficients are all given by

7⇣(3)
(3.9) K1wc = K2wc = K3wc = Nf ⇠02 ,
60

where ⇠0 = ~vf /2⇡kB Tc is the Cooper pair correlation length in the T = 0 limit. At

temperatures close to Tc the correlation length for spatial variations of the order parameter

is given by the GL coherence length,


s
K1 ⇠GL
(3.10) ⇠= = 1 ,
|↵(p, T )|
(1 T /Tc ) 2

1
where ⇠GL = (7⇣(3)/20) 2 ⇠0 in the weak-coupling theory for the gradient energies.

The strength of the quadratic Zeeman energy for spin-triplet pairing is given by

7⇣(3) Nf ( ~)2
(3.11) gzwc = ,
48⇡ 2 [(1 + F0a )kB Tc ]2

where is the nuclear gyromagnetic ratio for the 3 He nucleus and F0a is the exchange

interaction. The latter is ferromagnetic, varying from F0a = 0.723 at p = 0 bar to

F0a = 0.778 at melting pressure, p = 34 bar. Thus, combined with the large e↵ective

mass at high pressures the nuclear magnetic susceptibility is enhanced by an order of

magnitude relative to the Pauli susceptibility at the same density. This enhancement was
64

the basis for ferromagnetic spin-fluctuation exchange models for the superfluid transition

to spin-triplet pairing [77]. For convenience we include all relevant material parameters

as a function of pressure, with references to measured values, in Appendix 3.8.

3.2.2. Strong-Coupling Theory

A strong-coupling formulation of GL theory that accounts for the relative stability of the

bulk A- and B phases, and specifically the bulk A-B transition line, TAB (p) for pressures

above the polycritical point, p & pPCP was introduced in Ref. [73]. This strong-coupling

GL functional is defined by the fourth-order GL material parameters,

wc T sc
(3.12) i (p, T ) = i (p) + i (p) .
Tc

wc
The weak-coupling parameters, i (p), are obtained from the leading order contribution

to the Luttinger-Ward free-energy functional as an expansion in the small parameter


wc
Tc /Tf , where Tf ⇡ 1 K is the Fermi temperature. The i (p) are expressed in terms of

pressure-dependent material parameters as shown in Eq. 3.7, and can be calculated from

the material parameters provided in Table 3.2 of Appendix 3.8.

The next-to-leading order corrections to the weak-coupling GL functional enter as

corrections to the fourth-order weak-coupling material coefficients. These terms are of


sc wc 2
order i ⇡ i (T /Tf )hwi |T| i, where hwi |T|2 i is a weighted average of the square of
65

the scattering amplitude for binary collisions between quasiparticles on the Fermi sur-

face [78]. At high pressures, strong scattering of quasiparticles by long-lived spin fluc-

tuations largely compensates the small parameter T /Tf , resulting in substantial strong-

coupling corrections to the weak-coupling theory, and the stabilization of the AM state

as the A phase [79].

In the analysis of the stability of the vortex phases of 3 He-B we use improved re-

sults for the strong-coupling parameters based on a recent determination of the e↵ective

interactions and scattering amplitudes that account for the body of normal-state thermo-

dynamic and transport data on liquid 3 He over the full pressure range below the melting

pressure, as well as the heat capacity jumps for the bulk A and B phases at Tc (p) in zero

field [1]. The results of this analysis provide a quantitative theory for the thermodynamic

properties of the bulk A and B phases of superfluid 3 He at all pressures, including a quan-

titative determination of the bulk A-B transition line, TAB (p), for pressures above the

polycritial point, pPCP , as well as the temperature dependence of the free energy, entropy

and heat capacity at all temperatures below Tc . The strong-coupling corrections to the

-parameters obtained from microscopic theory [1], listed in Table 3.2 of Appendix 3.8,

reproduce the heat capacity jumps for the A and B transitions over the full pressure range.

In particular, the A phase correctly appears as a stable phase above the polycritical point

pPCP = 21.22 bar. However, in the standard formulation of the GL theory in which the

i parameters are evaluated at Tc , and thus treated as functions only of pressure, the A

phase is the only stable phase for all temperatures and pressures above pPCP , i.e. the

standard fourth-order GL theory fails to account for the bulk A-B transition at TAB (p).
66

In Ref. [73] the missing A-B transition line was traced to the omission of the tem-

perature dependence of the fourth-order parameters in the neighborhood of a triple

point. The latter is defined by the intersection of the second-order transition line given

by ↵(Tc , p) = 0, and the first-order boundary line separating the A- and B-phases given
1
by AB (TAB , p) ⌘ A B = 0, where A ⌘ 245 and B ⌘ 12 + 3 345
. At the PCP

we have TAB (pPCP ) = Tc (pPCP ). But, for p > pPCP the lines separate and we must retain

both the temperature and pressure dependencies of AB (T, p) to account for TAB (p)

in the vicinity of pPCP . The degeneracy between the A- and B-phases near pPCP is re-

solved by retaining the linear T dependence of the strong-coupling corrections to the

parameters. The suppression of the strong-coupling terms originates from the reduction

in phase space for quasiparticle scattering with decreasing temperatures, and is the basis

for the temperature scaling of the strong-coupling corrections in Eq. 3.12. The analysis

and predictions for the vortex phases of superfluid 3 He reported here are based on the

strong-coupling material parameters calculated and reported in Ref. [1], combined with

the known pressure-dependent material parameters, m⇤ , vf , Tc , and ⇠0 as listed in Table

3.2 in Appendix 3.8, and the temperature scaling in Eq. 3.12 that accounts for the re-

duction in strong-coupling e↵ects below Tc . The resulting bulk phase diagram predicted

by strong-coupling GL theory accounts remarkably well for the experimental A-B tran-

sition line, TAB (p), as shown in Fig. 3.1, as well as the heat capacity jumps of the bulk

A and B phases. We emphasize that the predictions of the relative stability of the A

and B phases by the strong-coupling GL functional is validated by microscopic calcula-

tions of TAB (p) [1] based on the formulation of the strong-coupling theory developed in

Refs. [78, 79, 80, 39].


67

3.3. Vortex States in Superfluid 3 He-B


For rotating equilibrium of superfluid 3 He-B the inter-vortex spacing for singly-quantized,
p 1/
2
axially symmetric vortices organized on a hexagonal lattice is d = / 3⌦ . For an an-

gular velocity of ⌦ = 1.7 rad/s the vortex unit cell dimension is d = 0.150 mm ⇡ 6.7⇥103 ⇠0

at p = 18bar. Thus, most of the vortex unit cell is occupied by a texture of the bulk B

phase,

(3.13) A↵i (r) = B R↵i [n̂, #] ei ,

where R↵i [n̂, #] is an orthogonal matrix that defines the relative angle of rotation, #,

about the local axis n̂, between the spin- and orbital coordinates of the Cooper pairs.

The texture, n̂(r), is determined by a competition of surface and bulk nuclear dipolar

enegies, modified by the pair-breaking and orienting e↵ects of the vortex flow and the

intrinsic vortex magnetization. These textural energies are treated perturbatively after

the vortex structure is calculated for a fixed choice of the relative orientation of the

spin and orbital coordinates of the Cooper pairs [71, 64]. In particular, we can neglect
p
the nuclear dipole energy for distances r < ⇠D = K1 /gD ' 15 µm ⇡ 6.7 ⇥ 102 ⇠0 at

p = 18 bar. Thus, we can choose a convenient computational cell dimension ⇠0 ⌧ dc ⌧ ⇠D

which allows a converged solution at distances well beyond the vortex core, but still at

distances well within the dipole coherence length. Thus we can omit the dipole energy

and work in a convenient spin- and orbital coordinate system. We use the basis of aligned

spin and orbital coordinates to determine the vortex structures and free energy of the

vortex states, and in the calculations reported here the computational cell dimension is

dc = 60⇠, where ⇠ is the temperature-dependent coherence length defined in Eq. 3.10.


68

10 1.0 10 10
0.9
0.8 5
5 0.7
5
0.6

y/⇠
y/⇠

y/⇠
0 0.5 0 0
0.4
5 0.3 5
0.2 5
0.1
10 0.0 10 10
10 5 0 5 10 10 5 0 5 10 10 5 0 5 10
x/⇠ x/⇠ x/⇠

Figure 3.2. Left panel: The axially symmetric B-phase vortex (“o-vortex”)
has a hard core with a node in (r). Center panel: The axially symmetric
A-core vortex (“v-vortex”) has a suppressed, but non-vanishing, condensate
density in the core which is predominantly the order parameter for the bulk
A phase. Right panel: The D-core vortex has a “double core” structure
that spontaneously breaks axial rotation symmetry. All plots are of the
condensate density, | (r)|2 ⌘ Tr AA† , in units of that for the bulk B-
phase, 2B ⌘ Tr AB A†B . The solutions of the GL equations for the o-
vortex, A-core and D-core vortices correspond to p = 10 bar and T = 0.25Tc ,
p = 34 bar and T = 0.75Tc , p = 20 bar and T = 0.55Tc , respectively.

3.3.1. Euler-Lagrange Equations

To determine equilibrium and metastable vortex phases we obtain stationary solutions

of the strong-coupling GL functional, F[A], defined by Eqs. 3.1-3.4. The equilibrium

and metastable states in zero field are solutions of the Euler-Lagrange equations of F[A]
69

defined by the functional gradient, G[A] ⌘ F/ A† = 0,

↵(T )A↵i + K1 r2 A↵i + (K2 + K3 )ri rj A↵j


⇥ ⇤
2 1 A↵i Tr AAT + 2 A↵i Tr AA† + T ⇤
3 (AA A )↵i

† ⇤ T

(3.14) + 4 (AA A)↵i + 5 (A A A)↵i = 0.

In zero magnetic field, at distances far from the core of a quantized vortex, |r| ⇠, the

order parameter approaches the bulk B phase order parameter with a global phase that

reflects the topological winding number of the vortex,

B
(3.15) A↵i (r) !p ↵i ei (r)
,
r!rc 3
H
where (r) is constrained by phase quantization, r ·d` = p 2⇡ with p 2 {0, ±1, ±2, . . .},

which we enforce with (r) = p imposed on the computational boundary, where is

the azimuthal angle in cylindrical coordinates defined with respect to phase singularity.

For external fields parallel to the axis of rotation, H = Hẑ, we must add a term

representing the Zeeman energy, gz H 2 ↵z Azi , to the left side of Eqs.3.14. In an external

magnetic field we must also modify the boundary condition to incorporate gap distortion

by the Zeeman energy on the bulk B phase order parameter. The boundary condition in

Eq. 3.15 is replaced by

⇥ ⇤
(3.16) A↵i (r) ! p1
3 ? ( ↵i ẑ↵ ẑi ) + k ẑ↵ ẑi ei (r)
,
r!rc
70

where the field-induced gap distortion of the order parameter is given by


s
12 H2
(3.17) ? = B 1+ 2
,
345 H0
s
2 12 + 345 H2
(3.18) k = B 1 ,
345 H02
p
where H0 ⌘ |↵(p, T )|/gz is the field scale at which the bulk B phase is strongly deformed

or destroyed.

In order to obtain stationary state solutions to the GL equations a simple method is

to find a solution of the discretized time-dependent GL equation [64],

@A↵i F
(3.19) = ⌘ G[A]↵i ,
@t A⇤↵i

which relaxes to a stationary state satisfying Eq. 3.14, i.e. G[A]↵i = 0. Here we use

the quasi-Newton, Limited-memory Broyden–Fletcher–Goldfarb–Shanno algorithm (L-

BFGS) [81, 82] to obtain stationary state solutions of G[A] = 0 that is far more efficient

than relaxation based on Eq. 3.19. Our implementation of the L-BFGS algorithm is

outlined in App. 3.9 where we also provide a benchmark comparison of the improvement

in rate of convergence to a solution of the GL equations using the L-BFGS algorithm

compared to relaxation 4.

In general there are multiple stationary-state solutions to Eq. 3.14. As a result con-

vergence to a steady-state solution can also be influenced by the initialization of the order

parameter. Thus, in addition to the boundary condition at the edge of the computational

cell, we use targeted initialization of the order parameter to find stationary states with

4See also Ref. [83] for implementations of gradient descent algorithms for solving the GL equations.
71

di↵erent symmetries. The free energy of the converged stationary solutions determines

the equilibrium phase. For example, to obtain a stationary solution for the D-core vortex,

either equilibrium or metastable, a non-axi-symmetric initialization of the order parame-

ter is used which converges to the targeted vortex efficiently. If the targeted vortex state

is not a local minimum then symmetry breaking at the initialization stage will not yield

a vortex with that broken symmetry.

Our analysis based on the strong-coupling free energy functional identifies the three

stationary state vortex solutions for 3 He-B in zero magnetic field, originally discussed by

Ohmi et al. [74] (o-vortex), Salomaa and Volovik [70] (A-core vortex) and Thuneberg [71]

(D-core vortex). Figure 3.2 illustrates the basic structure of these three vortices in terms of

their condensate densities. The o-vortex is “singular” with condensate density vanishing

at the vortex core center. The A-core vortex has a “superfluid core” with finite condensate

density in the core. The D-core vortex breaks rotational symmetry exhibiting a double

core structure, also with a finite condensate density.


Initialization & Soft Modes of the Order Parameter. The stationary “o-vortex” is

obtained by initialization of the order parameter as a singly-quantized local B-phase vortex


p
of the form A↵i (r) = p13 B tanh(|r|/ 2⇠) ↵i exp (i ). However, to target vortices with

lower symmetry we need to break additional symmetries, and it is useful to identify the

soft modes of the order parameter associated with relative spin-orbit rotation symmetry

of bulk 3 He-B.

In the absence of boundaries, magnetic fields, rotation and neglecting the nuclear

dipole energy the bulk B phase order parameter has a large degeneracy space associated

with relative spin-orbit rotations described by the rotation matrix R↵i [n̂, #], which defines
72

C++ ++ C+0 +0 C+ +
1.0

Amplitude
0.5

C0+ 0+ C00 00 C0 0
0

C + + C 0 0 C

Phase
0

Figure 3.3. Amplitudes and phases of the components of the o-vortex at


P = 10 bar and T = 0.25 Tc , shown on a square grid with x and y ranging
from [ 5 ⇠, +5 ⇠]. The computational grid was 60 ⇠ ⇥ 60 ⇠ with grid spacing
h = 0.1⇠. The o-vortex retains the maximal symmetry of the stationary
vortex solutions for the B-phase; axial rotation symmetry and time-reversal
symmetry are preserved. Amplitudes with N = 0 and N = 2 vanish by
symmetry. Thus, the o-vortex has vanishing condensate density in the core.

the orientation of the spin coordinates of the Cooper pairs relative to the orbital coordinate

axes.

For the analysis of the internal structure of vortex states in rotating 3 He-B, and their

relative energies, the spin-orbit rotational degeneracy is partially resolved by the vortex
73

flow. At distance scales ⇠ ⌧ r ⌧ ⇠D the spin-orbit rotational degeneracy is a soft mode

leading to some amplitudes of the order parameter developing long-range, power-law tails

/ 1/r, 1/r2 . The slow spatial variations of these modes for vortices in 3 He-B is discussed

in detail in Refs. [74, 84, 64, 85].

In our analysis we use the asymptotic behavior of the soft-modes to target specific

stationary vortex solutions. For the A-core vortex, we initialize the components, Axz and

Azx , to vary as 1/r and components, Axy and Ayx , to vary as 1/r2 for |r| > 5⇠. To target

the D-core vortex, we initialize by breaking axial symmetry by introducing a change in

sign between the Axz and Ayz amplitudes, and seed the cores of the amplitudes with 4⇡

phase winding (c.f. Eq. 3.25 in Sec. 3.3.2) by initializing Axz , Azx , Ayz , and Azy with

non-zero values in a small region of the core near r = (0, 0).

3.3.2. Axially Symmetric Vortex States in 3 He-B

For axially symmetric, singly-quantized vortices, the circulation of each vortex in the
H
asymptotic limit, |r| ⇠, r · d` = 2⇡, is satisfied by (r) = . The resulting

mass current and moduli of the amplitudes for all components of the order parameter

are axially symmetric. The simplest axially symmetric vortex is the local B-phase vortex

first discussed by Ohmi et al. [74]. The B phase of 3 He is invariant under joint spin and

orbital rotations. Thus, for axially symmetric vortices, or vortices with weakly broken

axial symmetry, it is instructive to represent the order parameter in the basis of angular
p
momentum eigenvectors, { µ |µ = 1, 0, +1}, where 0 = ẑ and ± = (x̂ ± iŷ)/ 2 [74].
µ ⌫⇤ µ⌫
These basis vectors satisfy the orthogonality relations, · = . We can transform a

p-quantized vortex from the spin-orbit aligned Cartesian basis to the angular momentum
74

basis by writing,

B
X
µ
(3.20) A↵i (r) = p ↵ [Aµ⌫ (r)] ⌫
i .
3 µ,⌫

where Aµ⌫ (r) ⌘ Cµ⌫ (r) eiN µ⌫ are the complex order parameter amplitudes in the angular

momentum basis, expressed in terms of amplitudes, Cµ⌫ (r), and phases, µ⌫ = Nµ⌫ . The

Nµ⌫ are integer winding numbers for the phase of the µ, ⌫ component. Asymptotically,

for a p-quantized vortex of the B-phase

B
(3.21) A↵i (r) !p ↵i eip .
|r|!1 3

A p-quantized B phase vortex that is also axially symmetric is an eigenstate of the

generator for axial rotations, i.e.

(3.22) Jz A↵i (r) = j~ A↵i (r) ,

with j 2 {0, ±1, ±2, . . .}. The total angular momentum projected along the axis of

symmetry, Jz = Lcm int int


z + Lz + Sz , is the sum of the operator for the center-of-mass orbital

angular momentum of the Cooper pairs, Lcm


z = i~@ , and the internal orbital and spin

angular momentum operators, Lint int


z and Sz . The latter yield,

(3.23) Lint
z

= ⌫~ ⌫
, Szint µ
= µ~ µ
.

The condition in Eq. 3.22 must also apply to the asymptotic limit in Eq. 3.21, which

requires j = p. Imposing the axial symmetry condition, Jz A↵i (r) = p A↵i (r), for any |r|

then fixes the phase of each (µ, ⌫) component, Nµ⌫ = p µ ⌫. Thus, the form of the
75

order parameter for a p-quantized, axially symmetric vortex becomes [74],

B
X ⇥ ⇤
µ
(3.24) A↵i (r) = p ↵ Cµ⌫ (r) ei(p µ ⌫) ⌫
i .
3 µ,⌫

For a singly-quantized (p = 1) B phase vortex we can organize the components into a

matrix labeled by the orbital and spin angular momentum indices,


0 1
i +i
B C++ e C+0 C+ e C
B C
(3.25) [Aµ⌫ ] = B
B C0+ C00 e +i
C0 e +2i C.
C
@ A
C + e+i C 0 e+2i C e+3i

In Fig. 3.3 we show the amplitude and phase structure of a stationary solution of

Eq. 3.14 for the most symmetric singly quantized vortex state in 3 He-B. This is the “o-

vortex”, or “normal-core vortex”, which is “singular” in the sense that all non-vanishing

components incur a phase winding, and therefore force these amplitudes to vanish as

|r| ! 0. This is clear from the results shown in Fig. 3.3 where the dominant components

are C+ , C00 and C +, all of which vanish as |r| ! 0. Each of these dominant amplitudes

have the same phase winding, + = 00 = + = , as shown in the corresponding phase

plots of Fig. 3.3. In addition, the o-vortex develops very small sub-dominant amplitudes,

C++ and C , with phase windings of N++ = 1 and N = +3, respectively, also shown

in Fig. 3.3.

A key observation regarding the o-vortex is that the two amplitudes with zero phase

winding, C0+ and C+0 , are identically zero. The amplitude C0+ represents the equal-spin,

chiral A phase with intrinsic angular momentum Jzint = +~ from the orbital state of the

Cooper pairs, while C+0 is the -phase, also with Jzint = +~ from the spin state of the
76

C++ ++ C+0 +0 C+ +
1.0

Amplitude
0.5

C0+ 0+ C00 00 C0 0
0

C + + C 0 0 C

Phase
0

Figure 3.4. Amplitudes and phases of the components of the A-core vortex
at P = 34 bar and T = 0.75 Tc , shown and computed on the same grid
as that in Fig. 3.3. The key amplitudes defining the A-core vortex are
the amplitudes with zero phase winding - the A phase C0+ and the spin-
polarized phase, C+0 . The amplitudes with winding number N = 2, C 0
and C0 are also important in terms of the relative stability between the
A-core and D-core vortex states, as discussed in Sec. 3.3.3.

Cooper pairs. Components with zero phase winding can support finite amplitudes in the

vortex core. This was the observation of Ref. [70], and the basis for the prediction of a

ferromagnetic vortex in which both amplitudes, C0+ and C+0 , are finite in the core. This

is the “A-core” vortex, which is a stationary solution of the GL equations (Eqs. 3.14).
77

At sufficiently high pressure the strong-coupling corrections that stabilize the bulk

A phase also stabilize the A-core vortex as the lowest energy vortex phase in 3 He-B.

As a result the A-core vortex has a “superfluid core”, with finite condensate density,

Tr AA† as shown in Fig. 3.2. Furthermore, the vortex circulation induces, via the

Barnett e↵ect,[86] a substantial spin polarization in the form of the phase, discussed in

more detail in Sec. 3.3.5.

Figure 3.4 shows a stationary solution of Eqs. 3.14 with axial symmetry which hosts

both the chiral A-phase (C0+ ) and phase (C+0 ) with non-zero amplitudes in the vortex

core. Note the large A-phase density, as well as the finite, but reduced, -phase density

in the core. Since there is no phase winding to suppress these amplitudes they grow to

values near the corresponding homogeneous bulk values of a superposition of confined A

and phases. Thus, the ratio of the two condensate densities in the A-core vortex is of

order |C+0 (0)|2 /|C0+ (0)|2 ⇡ 0.1 at p = 34 bar based on the strong-coupling enhancement

of the A phase as shown in Fig. 3.4.

3.3.3. Non-Axial Symmetric Vortex States in 3 He-B

Axial symmetry forces amplitudes with winding numbers N = 2, i.e. C0 and C 0 , to be

quadratically suppressed in the core as is shown in Fig. 3.4 for the A-core vortex (these

amplitudes are zero by symmetry for the o-vortex).

For doubly quantized vortices the quadratic suppression of the core amplitude, com-

bined with the cost in kinetic energy, generally leads to dissociation of doubly quantized

vortices into a pair of singly quantized vortices in order to recover lost condensation energy

for fixed total circulation.


78

C++ ++ C+0 +0 C+ +
1.0

Amplitude
0.5

C0+ 0+ C00 00 C0 0
0

C + + C 0 0 C

Phase
0

Figure 3.5. Amplitudes and phases of the components of the D-core vortex
at P = 20 bar and T = 0.55 Tc , shown and computed on the same grid as
that in Fig. 3.3. The key amplitudes and phases defining the D-core vortex
are those with N = 2 phase winding, C0 and C 0 . The D-core accomodates
the double phase winding by dissociation into two N = 1 vortices, allowing
the corresponding amplitudes to grow. This is important for the stability
of the D-core vortex relative to the A-core and o-vortex, as discussed in
Sec. 3.3.3. The dissociation of the N = 2 vortices is responsible for the
broken axial symmetry that is clearly shown in all the amplitudes and
phases.

Thus for the A-core vortex, if the amplitudes with N = 2 winding numbers were to

dissociate into a pair of N = 1 vortices then the result would be a gain in condensation

energy due to increased condensate amplitudes C0 and C 0 in the core.


79

10.0 10.0 2.5

7.5 7.5
1.5
5.0 5.0

2.5 2.5
0.5
y/⇠

y/⇠

y/⇠
0.0 0.0

2.5 2.5 0.5

5.0 5.0
1.5
7.5 7.5

10.0 10.0 2.5


10.0 7.5 5.0 2.5 0.0 2.5 5.0 7.5 10.0 10.0 7.5 5.0 2.5 0.0 2.5 5.0 7.5 10.0 1.0 0.5 0.0 0.5 1.0
x/⇠ x/⇠ x/⇠

Figure 3.6. Left: Axially symmetric current density of the A-core vortex
for p = 34 bar and T = 1.86 mK. The current is strongly suppressed to zero
by the growth of the A- and phases in the core for |r| . 2.5⇠. Center:
Anisotropic mass current flow field of the D-core vortex at P = 20bar,
T = 1.23 mK. Right: Expanded view of the current near the center of the
D-core showing the double vortex structure as the source of the anisotropic
current density. Currents are scaled in units of jc defined in Eq. 3.28.

The cost of dissociation is the potential reduction in core energy from the amplitudes with

zero phase winding. For the A-core vortex these amplitudes, C0+ and C+0 , with N = 0 are

favorable because of strong-coupling energies. Thus, there is a competition between a gain

in condensation energy by dissociation of the amplitudes with N = 2 winding numbers

and the loss in condensation energy of the N = 0 amplitudes favored by strong-coupling

and Zeeman energies. This competition is responsible for the stabilization of the D-core

vortex as the temperature is lowered below TV (p, H), where strong-coupling energies are

no longer sufficient to stabilize the axially symmetric A-core vortex, shown as the solid

(dashed) green phase boundary for zero field (H = 284 G) in Fig. 3.1.

At low pressures and low temperatures where strong coupling energies are relatively

small the A-core vortex is no longer competitive with the D-core vortex. Furthermore,
80

the o-vortex is never competitive with the D-core vortex, since forcing the N = 0, 2

components to vanish incurs too large a cost in condensation energy for the o-vortex

compared to the D-core vortex, even in weak-coupling theory.

The splitting of the N = 2 vortices into a pair of N = 1 vortices is shown clearly in

the plots of the phases 0 and 0 in Fig. 3.5, as is the growth in the amplitude for

these components compared to their suppressed values in the A-core vortex. What is

also clear is that the origin of the broken axial symmetry is the splitting of the N = 2

phase singularities. This splitting of the C0 and C 0 vortices along the y axis breaks

axial rotation symmetry, and generates a substantial uniaxial anistropy in the amplitudes

C0 and C 0 , as well as all other components. The connection between the broken axial

symmetry of the D-core vortex and the dissociation of the N = 2 vortices in C0 and

C 0 along the y axis is particularly evident in the mass current distribution discussed

below and shown in the right panel of Fig. 3.6, where the pair of dissociated mass current

vortices located at y ⇡ ±1.5 ⇠ dominate the internal structure of the D-core vortex mass

current distribution.

3.3.4. Mass Current Density

Galilean invariance in pure 3 He has important implications for the transformation of

velocities and mass currents in both normal and superfluid 3 He. In particular the order
u i2mu·r/~
parameter transforms as A↵i (r) ! A↵i (r) e under a Galilean boost with velocity

u. Thus, the phase of the order parameter undergoes a local gauge transformation, or

equivalently, vs ⌘ (~/2m)r#, transforms as a velocity field under a Galilean boost,


u
vs ! vs u. Galilean invariance also implies that the free energy density transforms as
u
f ! f j · u + O(u2 ) where j is the mass current density. For a boost from the rest
81

frame of the normal excitations, i.e. vn = 0, the gradient terms in the GL free energy
u
density transform as fgrad ! fgrad js · u + O(u2 ). Thus, by carrying out the boost

transformation we obtain the superfluid mass current density in the rest frame of the

excitations, expressed in terms of Cartesian components,

4m ⇥ ⇤
(3.26) js,i = = K1 A⇤↵j ri A↵j+K2 A⇤↵j rj A↵i+K3 A⇤↵i rj A↵j .
~

Far from the vortex core the phase gradient is small, |r#| ⌧ ⇡, or equivalently the

flow velocity is small compared to the maximum sustainable condensate velocity, i.e.

vs ⌧ vc = ~/⇠. Thus, the current reduces to its value in the London limit governed by

the local B phase order parameter in Eq. 3.21,

✓ ◆2
2m 2
(3.27) js = 2 K1 + 13 (K2 + K3 ) B vs , vs ⌧ v c .
~

The mass current recovers axial symmetry in the limit |r| ! 1, however, the anisotropic

corrections to axial flow decay slowly as 1/r2 . Equation 3.27 provides the characteristic

scale for the vortex mass currents in rotating 3 He-B,

(3.28) jc = 2(2m/~)2 (K1 + (K2 + K3 )/3) 2


B (~/⇠) .

Figure 3.6 shows the flow field for the mass current of both the A-core and D-core

vortices. The A-core vortex has an axial vortex flow that collapses and vanishes rapidly

in the zero phase-winding region of the A-phase and -phase core, |r| . 2.5⇠, as shown in

the left panel of Fig. 3.6. By contrast the broken axial symmetry of the D-core vortex is

evident in the center panel of Fig. 3.6. A zoomed region of the anisotropic core is shown
82

in the right panel which clearly shows the origin of the uniaxial anisotropy of the current

flow is the dissociation of the N = 2 vortices of C0 and C 0 into a pair of N = 1 vortices

at y ⇡ ±1.5⇠.

Another remarkable property of the D-core vortex, first reported in Ref. [64], is the

prediction of an axial current anomaly, i.e. the local pattern of currents flowing along the

axis of circulation, but with zero net mass transport and zero phase gradient along the

vortex axis 5. The z-axis current density can be expressed as

4m ⇥ ⇤
(3.29) js,z = = K2 A⇤↵j rj A↵z+K3 A⇤↵z rj A↵j
~
4m 2 ⇥ ⇤ i⌫ ⇤ ⇤ i⌫

(3.30) = = K2 Cµ⌫ e @⌫ Cµ0+K3 Cµ0 @⌫ Cµ⌫ e ,
3~ B
p
where @0 = @z and @± = (@x ⌥ i@y )/ 2. Figure 3.7 shows the z-axis current for the D-core

vortex for the same pressure and temperature as that for the vortex currents in Fig. 3.6,

also in units of jc . The axial current density spans an area of order A ⇡ 100 ⇠ 2 ⇡ 6.25 µm2 .

An idea for detection of the current anomaly along the z-axis is to inject electrons into

rotating 3 He-B from the outer, radial boundary. Electrons in 3 He-B form mesoscopic ions

of radius R ⇡ 1.5 nm [88]. The capture of these ions by D-core vortices should lead to

transport of the ions along the D-core vortex lines driven by the axial currents. Detection

of the ions by imaging on the top and bottom surfaces of the rotating vessel containing

5A similar axial current anomaly bound to disclination lines in the chiral phase of 3 He confined in a
cylindrical channel is reported in Ref. [87]
83

10
0.03

0.02
5
0.01

y/⇠
0 0.00

0.01
5
0.02

0.03
10
10 5 0 5 10
x/⇠

Figure 3.7. Axial mass current of the D-core vortex at p = 20 bar and
T = 0.55 Tc . The current density is scaled in units of jc .

superfluid 3 He-B, in much the same way in which vortices in rotating superfluid 4 He were

first imaged [89], would provide direct evidence of the axial mass currents 6.

3.3.5. Intrinsic Vortex Orbital and Spin Angular Momentum

To obtain the large scale structure of the vortex lattice for superfluid 3 He-B in equilib-

rium with a confining boundary rotating with a constant angular velocity, ⌦ , we must

transform the free energy functional to the frame co-rotating with the boundary poten-

tial. Equilibrium in a rotating frame is achieved by a Legendre transformation [90],

F0 = F J · ⌦ , where J = L + S is the total angular momentum of liquid 3 He, including

the orbital fluid angular momentum, L, and the nuclear spin angular momentum of 3 He.

The Legendre transformation leads to several energy scales associated with equilibrium

in a rotating frame. The dominant e↵ect of J · ⌦ is the entrainment of the normal

fluid into co-rotation with velocity vn = ⌦ ⇥ r. However, the superfluid velocity field
6Note that in contrast to the imaging of trapped ions in rotating 4 He the transport of ions in the D-core
phase of rotating 3 He-B would be under conditions of zero electric field.
84

is irrotational, and thus superfluid 3 He-B minimizes the kinetic energy and accomodates

co-rotation by the formation of a lattice of quantized vortices. A key observation is that

the orbital component of the Legendre transformation, L · ⌦, is achieved by introducing

⌦ ⇥ r). In particular, the gradient energy transformed


a gauge potential, a = (2m/~)(⌦

to the co-rotating frame can be written in terms of a kinetic energy density expressed

in terms of the co-variant derivative of the order parameter, r ! D = r ia, and a

coupling to the circulation of the gauge potential, r ⇥ a = (4m/~)⌦


⌦, to the intrinsic
0 0 0
orbital angular momentum of the Cooper pairs, fgrad = fkin + forbital ,

0
fkin = K1 (Dk A↵j )⇤ (Dk A↵j )

(3.31) + 1
K
2 s
[(Dj A↵j )⇤ (Dk A↵k ) + (Dk A↵j )⇤ (Dj A↵k )] ,

0 4m
(3.32) forb ⌘ Lorb · ⌦ = Ka "ijk = (A⇤↵i A↵j ) ⌦ k ,
~

where Ks = K2 + K3 and Ka = K2 K3 . In weak-coupling theory Kawc = 0, however,

particle-hole asymmetry and strong-coupling corrections give Ka ⇡ (kB Tc /Ef )2 K1wc , and

thus to an intrinsic orbital angular momentum of order Lorb = orb (n~/4) ( /Ef )2 , with

orb ⇠ O(ln(Ef /kB Tc )). The intrinsic orbital angular momentum is too weak to a↵ect the

relative stability of the vortex phases at typical rotation speeds.

However, there are perturbations that are important in understanding the vortex

structure and NMR signatures of the vortex phases of 3 He-B. In particular, in addition

to the nuclear dipole energy there is a contribution to the nuclear Zeeman energy that

is linear in the external field defined by the invariant, fz0 = m · H, where in terms of
85

Cartesian components,

(3.33) m = gz0 = AA† ↵


✏↵ ,

where m (s ⌘ m/ ) is the intrinsic nuclear magnetization (spin) density of the Cooper

pairs. The bulk B phase is time-reversal symmetric with sbulk ⌘ 0. Thus, the gyromag-

netic e↵ect observed in rotating 3 He-B is a manifestation of intrinsic spin polarization of

vortices, driven by vortex currents in the core region. This is a vortex manifestation of

the Barnett e↵ect [86], discussed in the context of vortices in the 3 P2 neutron superfluid

predicted to exist in the interiors of rotating neutron stars [91, 92].

The intrinsic magnetization (spin polarization) for axially symmetric vortices takes a

simple form when expressed in terms of amplitudes defined in the angular momentum

basis,

X
(3.34) m(r) = m0 |C+⌫ |2 ˆ + m? (r) .
|C ⌫ |2 ⌦

In addition to the axial component of the magnetization there is a transverse magneti-

zation density, m? (r), which integrates to zero for all stationary vortex states, both the

axially symmetric o-vortex and A-core vortex with m? = m? (r)r̂, as well as the axially

asymmetric D-core vortex. The magnitude of the intrinsic magnetization density is given

by [91, 92],

(3.35) m0 ⌘ gz0 2
B ⇡ n( ~) ln(Ef /kB Tc )( B /Ef )
2
.
86

While all spin-triplet vortices generate an intrinsic spin polarization, symmetry constraints

on the phase winding of the order parameter components that inhabit the vortex core, as

well as strong-coupling terms in the free energy functional that stabilize vortex core states

with zero phase winding, lead to vortex-core magnetic moments that reflect the symmetry

of the vortex core order parameter. In the case of the high-pressure phases of rotating
3
He-B the A-core vortices, which host the ferromagnetic phase in the core, possess a

substantial non-vanishing magnetization density in the cores. The D-core vortex phase

also has a substantial vortex magnetization, which also reflects the double-core structure

of that phase. The vortex magnetization density is shown for both phases in Fig. 3.8.
R
The total magnetic moment of the A- and D-core vortex phases, M = dr m(r) exhibits

a discontinuity at the first-order vortex phase transition. For example, the magnetization

per unit length (M/Lv ), per vortex jumps from MA /Lv = 2.95 m0 ⇠ 2 in the A-core phase

to MD /Lv = 5.36 m0 ⇠ 2 in the D-core phase at T = 2.0 mK and p = 15.0 bar.

The direction of the vortex magnetization is selected by the angular velocity, i.e.
ˆ . As a result the linear Zeeman energy, f 0 =
mV = m(r)⌦ ˆ · H is the origin of the
m(r)⌦
z

gyromagnetic e↵ect observed in the NMR spectrum for the phases of rotating 3 He-B.

3.4. Magnetic Susceptibility

At sufficiently low magnetic fields the magnetization is determined by the nuclear

Zeeman energy,

Z
(3.36) FZeeman = 1
2
d 3 r H↵ ↵ (r)H ,
87

0.175
10 mD (x, 0)
0.150 mD (0, y)
y/⇠
0
mA (0, y)
0.125
10
m(r)/m0

10 0 10
0.100 x/⇠
10
0.075

y/⇠
0
0.050
10
10 0 10
0.025 x/⇠

0.000
20 15 10 5 0 5 10 15 20
x, y[⇠]

Figure 3.8. Magnetization profiles, mA and mD , for the A- and D-core


vortices, respectively. Insets: density plots of the same. For the A-core
phase: p = 34 bar T = 0.75 Tc = 1.86 mK. For the D-core phase: p = 20
bar T = 0.55 Tc = 1.23 mK.

evaluated with the zero-field order parameter, i.e. neglecting order parameter distortion

by the external field. For fields H k z the corresponding local magnetic susceptibility is

X
(3.37) zz (r)/ N =1 2gz |Azi (r)|2 .
i
88

0.80 A
(0, y)
D
(x, 0)
0.75
D
(0, y)
0.70
zz (r)/ N

0.65

0.60 10 10 0.8

0.55
y/⇠

y/⇠
0 0 0.6
0.50
-10 -10 0.4
-10 0 10 -10 0 10
0.45 x/⇠ x/⇠

20 15 10 5 0 5 10 15 20
x, y[⇠]

Figure 3.9. Susceptibility profiles, A zz and


D
zz , for the A- and D-core
vortices, respectively. Calculations are for the zero-field order parameters
at p = 34 bar just to the right and left of the zero-field transition line: A zz
( D
zz ) is evaluated at T = 1.80 mK (T = 1.79 mK). Insets: density plots of
the same.

Figure 3.9 shows our results for the local susceptibilities of the A- and D-core vortices.

The D-core vortex has the larger susceptibility, and thus we expect that the equilibrium

vortex transition line to shift to higher temperatures with the application of a weak

magetic field. This is indeed what we find from self-consistent solutions of the the GL

equations when we include the Zeeman energy in Eq. 3.3. Figure 3.10 shows the evolution
89

of the equilibrium vortex transition temperature with field, TV (p, H), for p = 34 bar.

The initial increase of TV with field is indicative of susceptibilities for the A- and D-core

vortices. However, at fields H & 40 G the transition temperature reaches an maximum,

then decreases with increasing field, such that TV (p, H = 284 G) = 1.755 mK < TV (p, H =

0) = 1.787 mK. The increase in TV relative to the zero-field transition for H & 60 mK

results from distortion on the vortex-core order parameters by the field, which dominates

the Zeeman term even at relatively low fields due to the near degeneracy of the two vortex

phases. This leads to the equilibrium vortex phase transition line, TV (p, H), for H = 284 G

shown in Fig. 3.1. The equilibrium transition line, TV (p, H), as well as the supercooling

transition line, TV⇤ (p, H), are reported for fields H||⌦
⌦||ẑ, and for the background B-phase

order parameter given in Eq. 3.16. Our results neglect the dipolar interaction within

the computational cell, ⇠ ⌧ dc ⌧ ⇠D , but include the e↵ects of vortex counterflow and

field-induced gap distortion. The weak nuclear dipole energy of 3 He-B confined in the

cylindrical experimental cell used in the rotating 3 He-B experiments reported in Ref. [25]

introduces a large-scale texture of the background B-phase that varies on the scale of the

cell radius, R ⇡ 2.5 mm.7 A discussion of the e↵ects of non-axial magnetic fields, as well

the possibility of weak inhomogenous broadening from large scale textural e↵ects, on the

vortex phases and p T H phase diagram is outside the scope of this article.

We note that Kasamatsu et al. recently published a report on the e↵ects of non-axial

magnetic fields on the structure of vortices in rotating 3 He-B [6]. They use the set of

7In particular, the anisotropy axis defining the B-phase order parameter in Eq. 3.13 aligns along H||z
in the center of the cell, but tilts away from the z-axis at an angle of (r), with (r) ' 1 r, with
1 . 4 r/R for pressures p & 20 bar and rotation speeds ⌦ . 2 rad/s. The texture leads to the transverse

NMR shift and the spectrum of spin-wave bound states. The e↵ects of vortex counter flow confined in
the vortex cores, as well as field-induced gap distortion tune the slope, and thus are observable in the
NMR spectrum [25].
90

strong-coupling GL material parameters obtained from analysis of several experiments by

Choi et al. [93]. However, their calculations are based on the standard GL free energy

functional [64]. This limits their analysis of relative stability of vortex phases in 3 He-B to

pressures below the polycritical point pressure, pPCP = 21.22 bar, and temperatures very

close to Tc , thus precluding an analysis of the stability of phases over the experimentally

relevant region of the p T H phase diagram. Our analysis, based on the strong-

coupling GL theory discussed in Sec. 3.2, allows us to explore the entire pressure range,

and specifically the phase diagram above the polycritical point pressure and temperatures

below the bulk A-B transition, which is the region most relevant to the phases and phase

transitions observed in rotating 3 He-B.

3.5. Equilibrium & Metastability Transitions

The experimental transition between the two distinct vortex phases of rotating 3 He-B

is hysteretic as shown in Fig. 1 of Ref. [5]. The vortex phase transition on cooling occurs

at much lower temperature than the phase transition on warming. This is indicated on

the pressure-temperature phase diagram for p = 29.3 bar the transition on cooling occurs

at TV⇤ = 1.43 mK while the transition on warming occurs at a higher temperature which

we estimate to be TV = 1.81 mK. The latter was identified as the temperature at which

the NMR satellite frequency splitting measured on warming merges with that measured

on cooling. There is some uncertainty in this value because both A-core and D-core

vortices are local minima of the free energy functional. Thus, on warming the heat flux

of quasiparticles may heat the vortex cores and prematurely convert some D-core vortices

to A-core vortices. Thus, a smooth extrapolation of the NMR splitting on warming yields
91

1.80

1.79
A-Core
1.78
T [mK]
V

1.77
D-Core
1.76

1.75
0 50 100 150 200 250 300
H [G]

Figure 3.10. Field evolution of the equilibrium vortex transition tempera-


ture at p = 34 bar.

TV ⇡ 1.85 mK, also indicated in Fig. 3.1. This is the only data we found in the literature

for the transition on warming. The data for the transitions on cooling for all reported

pressures was obtained from Fig. 2 of Ref. [5]. All data reported in Fig. 3.1 of this report

was converted from the Helsinki temperature scale to the widely accepted Greywall scale

according to TGreywall = 0.89 THelsinki [3]. The transitions on cooling all exhibit a sharp

drop in the NMR frequency at the same temperature independent of rotation speed. There
92

is no further supercooling, indicating that TV⇤ is a global instability below which there is

only one phase that is a local minimum of the free energy.

The theoretical results we report for the phase diagram in Fig. 3.1 are based on precise

numerical solutions of the strong-coupling GL equations for the vortex phases of rotating
3
He-B. The experimental transition on warming at p = 29.3 bar and H = 284 G is in

close agreement with our determination of the equilibrium transition line, TV (p, H), at

that pressure and field. We identify the warming transition as the equilibrium vortex

phase transition, i.e. point in the (p, T ) plane where the free energies of the two phases

are equal. This interpretation is based on our calculations of the free energies of the high-

temperature, high-pressure A-core phase and the low-temperature, low-pressure D-core

phase. In particular, the equilibrium transition line calculated as the locus of points where

the A-core and D-core free energies are equal is shown in Fig. 3.1 for zero field as the solid

green line. This transition line terminates on the bulk transition line at a triple point

[pvc , Tc (pvc )]. Thus, there is a window within the B phase where the A-core vortex phase

is the equilibrium phase even in zero field, with the A phase and phase inhabiting the

cores of vortices within the A-core phase. The A phase is able to grow within the B-phase

vortex core because of the suppression of the B-phase amplitudes with winding number

N = 1: C00 , C+ and C + and the absence of any suppression for the N = 0 amplitudes:

C0+ and C+0 . Thus, with strong-coupling support for the A-phase the A-core vortex is

stabilized at sufficiently high pressure and high temperature in the region shaded in green

in Fig 3.1. Also shown is the equilibrium region of the A-core vortex phase for the field of

H = 284 G ẑ. Note that the equilibrium region of the A-core phase is extended to lower

temperatures (c.f. Fig. 3.10) and pressures within the range, TV (p, H) < T < TAB (p, H),
93

as shown by the dashed green line in Fig. 3.1. Our analysis also shows that the region

between the transition at TV (p, H) and the transition at TV⇤ (p, H) corresponds to the

region in which the high-temperature A-core vortex phase is a metastable local minimum

of the free energy, but is not the global minimum. Thus, the A-core phase supercools to

the lower temperature, TV⇤ , below which the high-temperature A-core phase is globally

unstable to the D-core phase. Indeed, the observed transitions on cooling for H = 284 G,

over the pressure range 20 bar . p  34 bar, agree well with our theoretically determined

metastability transition, TV⇤ (p, H), at which the A-core vortex phase is globally unstable

to the D-core vortex phase.

The supercooling transition at TV⇤ (p, H) shown as the dashed purple line in Fig. 3.1,

and the much larger region of metastability of the A-core vortex phase (shaded in pink),

was obtained by starting at high pressure and high temperature in the region of global

stability of the A-core vortex phase, then lowering temperature slightly below TV (p, H),

where the D-core vortex is the global miniumum, and initializing the order parameter with

the higher temperature A-core order parameter field plus a small admixture (“seed”) of

the D-core order parameter, i.e. Ainit (p, Tnew ) = AA-core (p, Tlast ) + ✏ AD-core (p, Tnew ), where

✏ ⌧ 1. Throughout the region bounded by TV (p, H) and TV⇤ (p, H) (shown in pink) the

vortex initialized with the D-core perturbation returned to the axially symmetric A-core

phase. The supercooling transition, TV⇤ (p, H), was the locus of points where the A-core

was globally unstable. We note that results for the supercooling transition require a fine

computational grid. For a coarse grid of h = 0.5⇠ the supercooling transtion is lower

than that shown in Fig. 3.1, but converges to the reported transition line for h . 0.15⇠.

The phase transition lines shown in Fig. 3.1 were obtained on a 60⇠ ⇥ 60⇠ computational
94

grid with grid spacing h = 0.1⇠. Our numerical annealing procedure used to identify the

region of metastability of the A-core phase agrees remarkably well with the experimental

results for the transition obtained on cooling, both in the magnitude of the supercooling at

pressures above pcv , as well as the rapid cross over in slope of TV⇤ (p, H) with pressure at the

lower pressures approaching pcv . However, our region of metastability does not extend as

low in pressure as the experimentally reported transitions on cooling. Our interpretation

of the latter is that below pcv strong-coupling energies are never able to stabilize the A

phase in the vortex core, without assistance from the Zeeman energy. This results in the

termination of the supercooling line on the equilibrium A-core vortex phase boundary

at a pressure near pcv . We are not able to resolve the origin of the discrepancy in the

minimum pressure for the metastable A-core phase within the strong-coupling GL theory.

Such a resolution may require new experiments under rotation with pressure sweeps, or

perhaps implementation of the full quasiclassical strong-coupling free energy functional

extended to inhomogeneous phases.

3.6. Summary and Outlook


The recent development of a strong-coupling Ginzburg-Landau theory that accounts

for the relative stability of the bulk A and B phases has provided the first opportunity

to examine the relative stability of the vortex phases discovered in rotating 3 He-B and

to predict, based on known material properties of superfluid 3 He over the full pressure

range, the equilibrium and metastable vortex phase transitions. We are able to verify

the local and global stability of all the stationary solutions to the strong-coupling GL

theory over the full (p, T ) plane. Only the A-core and D-core phases are global minima

anywhere in the (p, T ) plane. The results we report provide strong theoretical support for
95

the identification of the experimentally observed phase transitions as the equilibrium and

supercooled phase transitions between the high temperature A-core vortex phase with

broken time-reversal and mirror symmetries (proposed by Salomaa and Volovik [70]),

and the low temperature, low pressure D-core vortex phase with broken axial symmetry

(proposed by Thuneberg [71]). Furthermore, both of these transitions are driven by

the decrease in strong coupling energies at sufficiently low pressures and temperatures

defined by the metastability line TV⇤ (p, H). In addition, the broken rotational symmetry

of the D-core vortex is identified with the instability of the components within the core

with 4⇡ phase winding. Once strong-coupling energies are suppressed by sufficiently low

temperature or pressure the doubly quantized vortices dissociate to gain condensation

energy, and as a result break axial symmetry.

We conclude with the two forward looking observations. First, the success of the

strong coupling GL theory, evident by the results for the vortex phase diagram, provides

a theoretical tool for studying a wide range of problems involving inhomogeneous phases

with complex symmetry breaking and/or novel topological defects, in the strong-coupling

limit, that were not previously accessible. A recent example is the analysis of the exper-

imentally measured Bosonic collective mode frequencies (“Higgs masses”) of superfluid


3
He-B using a time-dependent extension of the strong-coupling GL theory in Ref. [94],

which provided consistent experimental results for the strength of the f-wave pairing in-

teraction in superfluid 3 He over the full pressure range [95], a material parameter that is

important for understanding ground-states and excitations of superfluid 3 He at high pres-

sures and high magnetic fields. Secondly, the strong-coupling GL theory is supported by

the microscopic strong-coupling pairing theory based on leading order corrections to the
96

weak-coupling BCS theory originating from binary collision scattering between fermionic

quasiparticles of the normal phase of liquid 3 He [1]. Further development of a quantitative

microscopic strong-coupling pairing theory to inhomogeneous, non-equilibrium states is

well within reach.

3.7. Acknowledgements
A preliminary report of these results was presented at the International Conference on

Quantum Fluids and Solids (QFS2018), Tokyo, Japan in July 2018. We thank Wei-Ting

Lin for discussions on efficient numerical approaches to solving multi-component Euler-

Lagrange PDEs, and Erkki Thuneberg for detailed comments on a preliminary version of

this manuscript. The research was supported by the National Science Foundation (Grant

DMR-1508730).

3.8. Appendix: Material Parameters

The following tables summarize the pressure dependent material parameters that de-

termine the properties of the superfluid phases in strong-coupling theory.

n sc sc sc sc sc
1 2 3 4 5
0 9.849 ⇥ 10 3 4.193 ⇥ 10 2 1.322 ⇥ 10 2 4.747 ⇥ 10 3 8.987 ⇥ 10 2

1 5.043 ⇥ 10 2 1.177 ⇥ 10 1 5.428 ⇥ 10 2 3.788 ⇥ 10 1 6.925 ⇥ 10 1

2 2.205 ⇥ 10 2 4.322 ⇥ 10 2 9.559 ⇥ 10 2 1.774 ⇥ 10 1 8.761 ⇥ 10 1

3 2.557 ⇥ 10 2 8.793 ⇥ 10 2 6.419 ⇥ 10 2 1.735 ⇥ 10 1 5.929 ⇥ 10 1

4 5.023 ⇥ 10 2 8.598 ⇥ 10 2 9.310 ⇥ 10 3 1.878 ⇥ 10 1 2.904 ⇥ 10 2

5 2.769 ⇥ 10 2 3.639 ⇥ 10 2 1.862 ⇥ 10 2 1.522 ⇥ 10 1 8.870 ⇥ 10 2

Table 3.1. Coefficients of a polynomial fit to the strong-coupling param-


P (i)
eters from Ref. [1] of the form isc = n an pn .
97

p[bar] n [nm 3 ] m⇤ /m F0a Tc [mK] vf [m/s] ⇠0 [nm] sc


1
sc
2
sc
3
sc
4
sc
5
0.0 16.28 2.80 -0.7226 0.929 59.03 77.21 -0.0098 -0.0419 -0.0132 -0.0047 -0.0899
2.0 17.41 3.05 -0.7317 1.181 55.41 57.04 -0.0127 -0.0490 -0.0161 -0.0276 -0.1277
4.0 18.21 3.27 -0.7392 1.388 52.36 45.85 -0.0155 -0.0562 -0.0184 -0.0514 -0.1602
6.0 18.85 3.48 -0.7453 1.560 49.77 38.77 -0.0181 -0.0636 -0.0202 -0.0760 -0.1880
8.0 19.34 3.68 -0.7503 1.705 47.56 33.91 -0.0207 -0.0711 -0.0216 -0.1010 -0.2119
10.0 19.75 3.86 -0.7544 1.828 45.66 30.37 -0.0231 -0.0786 -0.0226 -0.1260 -0.2324
12.0 20.16 4.03 -0.7580 1.934 44.00 27.66 -0.0254 -0.0861 -0.0233 -0.1508 -0.2503
14.0 20.60 4.20 -0.7610 2.026 42.51 25.51 -0.0275 -0.0936 -0.0239 -0.1751 -0.2660
16.0 21.01 4.37 -0.7637 2.106 41.17 23.76 -0.0295 -0.1011 -0.0243 -0.1985 -0.2801
18.0 21.44 4.53 -0.7661 2.177 39.92 22.29 -0.0314 -0.1086 -0.0247 -0.2208 -0.2930
20.0 21.79 4.70 -0.7684 2.239 38.74 21.03 -0.0330 -0.1160 -0.0249 -0.2419 -0.3051
22.0 22.96 4.86 -0.7705 2.293 37.61 19.94 -0.0345 -0.1233 -0.0252 -0.2614 -0.3167
24.0 22.36 5.02 -0.7725 2.339 36.53 18.99 -0.0358 -0.1306 -0.0255 -0.2795 -0.3280
26.0 22.54 5.18 -0.7743 2.378 35.50 18.15 -0.0370 -0.1378 -0.0258 -0.2961 -0.3392
28.0 22.71 5.34 -0.7758 2.411 34.53 17.41 -0.0381 -0.1448 -0.0262 -0.3114 -0.3502
30.0 22.90 5.50 -0.7769 2.438 33.63 16.77 -0.0391 -0.1517 -0.0265 -0.3255 -0.3611
32.0 23.22 5.66 -0.7775 2.463 32.85 16.22 -0.0402 -0.1583 -0.0267 -0.3388 -0.3717
34.0 23.87 5.82 -0.7775 2.486 32.23 15.76 -0.0413 -0.1645 -0.0268 -0.3518 -0.3815

Table 3.2. Material parameters for 3 He vs. pressure, with the particle
density n = kf3 /3⇡ 2 from Ref. [2], the e↵ective mass, m⇤ , and Tc from
Ref. [3], the exchange interaction, F0a , is from Ref. [4], the Fermi velocity,
vf = ~kf /m⇤ , calculated from the Fermi wavelength, kf , and the coherence
length is ⇠0 = ~vf /2⇡ kB Tc . The strong-coupling parameters, isc , in units
of | 1wc |, are from Ref. [1].

3.9. Appendix: Numerical Methods

The search for stationary states of the Ginzburg-Landau functional leads to the Euler-

Lagrange Equations (Eqs 3.14), which are coupled, non-linear partial di↵erential equations

(PDEs) for the 18 components of the 3 He order parameter. The method of relaxation

based on the discretized version of Eq. 3.19 to improve the approximate solution at each

step along the gradient direction until one reaches the steady-state solution is generally

inefficient. Instead, we employ an efficient numerical method developed to solve the multi-

component field equations, e.g. the order parameter for topological defects in superfluid
3
He. The method is based on the L-BFGS optimization algorithm [82] summarized below.
98

L-BFGS
1
Relaxation
0
log(F/F0)

5
0 500 1000 1500 2000
Iterations

Figure 3.11. Rate of convergence of the iterative solution of the GL equa-


tions expressed in terms of the Free energy at each iteration, k, for the
o-vortex at P = 10 bar, T = 0.25 Tc normalized by the bulk B phase free
energy integrated over the same volume. The L-BFGS algorithm is shown
in green, while the relaxation algorithm is shown in red.

In Newton’s method, we solve the equation

(3.38) xk+1 = xk + ↵k pk

where ↵k ⌘ ↵ represents a fixed step size at each iteration labeled by k, and pk = Hk 1 G k

is the search direction where Gk is the functional gradient defined in Eq. 3.14. Hereafter
99

we follow standard notation [82] and denote the inverse Hessian simply by Hk . We

implement this algorithm by storing the order parameter, A↵i (x, y), as a four-dimensional

(↵,i,x,y) array of complex numbers represented by xk at iteration k. At each iteration,

the order parameter is updated along with the step size and search direction. Storing the

exact inverse Hessian, Hk , requires calculating a matrix of N ⇥N second derivatives which

is computationally expensive. Instead, we use the L-BFGS quasi-Newton minimization

algorithm. This requires us to solve Eq. 3.38 to determine a step size, ↵k , that minimizes

the function f (↵k p̂ + xk ) at each iteration, where the direction p̂ is constructed from the

approximation to the inverse Hessian Hk , using Gk and xk . We require the inverse Hessian

to be symmetric and positive definite. To determine Hk+1 we solve the minimization

problem

(3.39) min ||Hk H||, H = H † , Hyk = sk ,


H

where sk = xk+1 xk , yk = Gk+1 Gk , where Gk is the functional gradient of the GL func-

tional at iterate k, Gk = Fk / A⇤↵i,k . The unique solution to this minimization problem is

obtained by re-writing the minimization problem in terms of a weighted Frobenius norm,

which transforms the minimization problem to a new basis under a unitary transforma-

tion. The Frobenius norm can then be calculated explicitly and minimized. Transforming

back to the original basis we obtain the solution

(3.40) Hk+1 = (1 ⇢k sk yk† )Hk (1 ⇢k yk s†k ) + ⇢k sk s†k ,


100

where ⇢k = 1/(yk† sk ). Eq. 3.40 is known as the BFGS update [82], and is an approximation

to the inverse Hessian Hk+1 given an initial inverse Hessian Hk . Thus, we now solve

Eq. 3.38 with a search direction given by pk = Hk Gk which is calculated in terms of

inner products of the form hyk |Gk i, hsk |Gk i. This makes the solution of Eq. 3.14 straight-

forward with

(3.41) sk = xk+1 xk , yk = Gk+1 Gk .

We initialize the inverse Hessian with H0 = 1, then update according to

s†k 1 yk1
(3.42) Hk = †
.
yk 1 yk 1

This is an approximation to the inverse Hessian matrix along the most recent search

direction. For the L-BFGS update at iteration k we have the current iterate as xk and

we store a limited memory set of vector pairs {si , yi } for i = k m, .., k 1. Thus, by

choosing an initial approximate inverse Hessian Hk0 we obtain by repeated iteration of

Eq. 3.40 the L-BFGS algorithm [82],

(3.43) Hk = (Vk† 1 · · · Vk† 0


m ) Hk (Vk m · · · Vk 1 )


+ ⇢k m (Vk 1 · · · Vk† †
m+1 )sk m sk m (Vk m+1 · · · Vk 1 )


+ ⇢k m+1 (Vk 1 · · · Vk† †
m+2 )sk m+1 sk m+1 (Vk m+2 · · · Vk 1 )

+ . . . + ⇢k 1 sk 1 s†k 1 .
101

The arrays sk , yk which encode the order parameter and functional gradient, are stored

as five-dimensional complex arrays where one component of the array is a memory index

and the other four components represent the orbital, spin and spatial degrees of freedom

in the x y plane. The L-BFGS algorithm is used to calculate the stationary states by

solving Eq. 3.14 for the full (p, T ) plane. In Fig. 3.11 we compare numerical relaxation

with the rate of convergence of the L-BFGS algorithm for the axially symmetric o-vortex.

The performance of the L-BFGS algorithm is essential in being able to calculate the

equilibrium and metastable phase diagram on reasonable timescales.

3.10. Appendix: Benchmarking the GL Solver

We tested our code against others by comparing our results for the free energies of

the A-core and D-core vortex states with those reported in Ref. [6] based on their choice

for the GL parameters. Figure 3.12 shows results based on our GL solver using GL

parameter set I of Ref. [6] for p = 34 bar. Our result is in excellent agreement with the

result reported by Kasamatsu et al. in their Fig. 5(d) for the same GL parameters,

including the crossing field of H ⇡ 100 G. N.B. While these are local minima of the GL

functional for this parameter set, they do not represent realized solutions at this pressure

because this set of GL parameters does not account for the relative stability of the

bulk A and B phases for pressures above pPCP = 21.22 bar. Nevertheless, the comparison

provides a benchmark and additional confidence in our GL solver and numerical results.
102

0.5
D-Core

0.0
A-Core

0.5
F̃V

1.0

1.5

2.0
0 50 100 150 200 250 300
H(G)

Figure 3.12. Free energies for the A-core and D-core vortex states versus
magnetic field calculated using our GL solver for the set I GL parameters
at p = 34 bar. The core contribution to the free energies are calculated as
in Ref.[6] by subtracting the bulk and hydrodynamic contributions to the
B phase energy, and normalizing the result in units of the bulk B phase
energy at each field.
103

CHAPTER 4

Half-Quantum Vortices in Superfluid 3 He Confined in Nafen

Based on a strong-coupling Ginzburg-Landau (GL) theory that accounts for the rel-

ative stability of the confined superfluid phases of 3 He infused in Nafen, we report the-

oretical calculations for the structure of half-quantum vortices (HQVs) in these phases.

We identify two di↵erent ground state vortex phases within the polar and polar-distorted

chiral A-phase. In the polar phase, a HQV pair with a polar core is discovered. The core

and asymptotic structure of a polar HQV is investigated numerically and analytically, and

is determined to have a superfluid, spin-polarized ferromagnetic phase core [96]. The

HQV pairs are also calculated in the presence of an axial magnetic field, H = 0 ⌦,
370G⌦

and no quantitative di↵erence is observed. The evolution of a polar HQV pair is then

calculated in a transverse magnetic field which weakens the stability of the vortex until

the pair collapses into a singly-quantized vortex for H > 1000Gx̂. The zero-field trans-

verse magnetic susceptibilities are calculated and we find that in the vortex cores the

vortices are equally polarizable along either applied field direction. The evolution of the

Barnett induced magnetization as a function of inter-vortex spacing is also analyzed and

we find it vanishes when the HQV pair converts into a singly-quantized vortex. The free

energy as a function of inter-vortex separation is calculated which indicates that the pairs

of HQVs are at lowest energy when the separation is largest. The singly-quantized vor-

tex is higher energy than the pair of HQVs which indicates that the HQVs are indeed a
104

local minima of the free energy, and not a metastable state. We then lower the temper-

ature transitioning from the polar phase into the polar-distorted chiral A-phase, which

hosts a newly discovered polar-distorted chiral HQV. We first analyze the chiral phase

and discover spontaneous supercurrents arising from the intrinsic coupling between the

di↵erent orbital degrees of freedom. We then calculate the structure of the chiral vortex

in this phase and find that this vortex is characterized by having both a chiral and polar

component in its core, and also a spontaneous supercurrent along the polar axis. These

pairs of HQVs are the equilibrium states in the confined phases which are stabilized by

infusing superfluid 3 He in a highly anisotropic material known as Nafen aerogels. Super-

fluid 3 He in confined phases has been recently summarized in great detail [97]. Here, the

confined interaction is modeled by constructing an array of nematic line impurities on a

lattice in which the HQVs interact. We also predict the regions of energetic stability for

two di↵erent ground state vortex phases within these confined phases. We investigate a

pair of polar HQVs that are stabilized in the polar phase, and a newly discovered pair

of polar-distorted chiral HQVs in the polar-distorted chiral A-phase. We lastly calcu-

late the confined superfluid phase diagram in a magnetic field along the rotation axis,

H=0 370G ⌦ , and obtain excellent agreement with the experimentally reported phase

transitions observed in the equal-spin pairing phases of 3He. Our results provide a defini-

tive identification of the topological defects order parameter, free energy, intrinsic orbital

angular momentum and magnetization, magnetic susceptibility, and mass currents.

Quantized vortices are the topological defects of superfluids and superconductors that

reflect the non-trivial topology of the degeneracy space of the order parameter manifold.

In spin-singlet superfluids and conventional superconductors, the order parameter has


105

a single degree of freedom and is characterized by a scalar order parameter. In 3 He,

BCS condensation arises from Cooper pairs of 3 He spin-triplet quasiparticles, thus the

order parameter has 2(2l + 1)(2s + 1) = 18 degrees of freedom. This complicated order

parameter structure allows for the stabilization of many di↵erent kinds of vortices in the

various phases of 3 He. In this paper, we will discuss quantized vortices carrying half a

unit of circulation, known as the half-quantum vortex (HQV), in the equal-spin pairing

(ESP) confined phases of 3 He. The two ESP phases in confined superfluid 3 He are the

polar phase, and the polar-distorted chiral A phase. These confined phases were both

discovered experimentally when infusing 3 He into aerogels [7]. The polar phase is an ESP

state in which the orbital momentum is aligned along the strands of the nematic aerogel,

ẑ. A complete detailed review of superfluid 3 He in aerogel is provided recently [98]. The

polar phase hosts a polar HQV which is characterized by an equal superposition of Cooper

pairs with equal and opposite spin projections [99]. The polar-distorted chiral A phase has

a dominant orbital component along ẑ and a sub-dominant component along x̂, where the

chiral axis is along ŷ. We find the polar-distorted chiral HQV as the equilibrium vortex

state in this phase.

Experimental realization of the confined superfluid phases of 3 He, and evidence of

HQVs in the polar phase of topological superfluid 3 He was strong motivation for this

work [100, 7, 28, 101]. The vortices were generated by rotating 3 He in a cyrostat and

also under zero rotation by the Kibble-Zurek mechanism. There has been considerable

detailed analysis on the formation of topological defects from the Kibble-Zurek mecha-

nism [102, 103, 104, 105]. The HQV is an exotic topological defect in superfluids and

superconductors since it carries a half quantum of circulation, and additionally can host
106

Majorana zero modes bound to its cores [106]. The polar phase is stabilized when 3 He is

confined to Nafen aerogel impurity strands, known as nematic aerogels [107, 7, 108].

Stability of the polar phase in silica aerogels has also been analyzed in great detail

[109, 110, 111, 112]. Various theoretical models on 3 He in anisotropic mediums and aero-

gel have been studied previously 3 He [113, 114, 1]. There has also been considerable theo-

retical work done in the context of weak-coupling GL theory [115, 116, 117, 118, 119].

Additionally, NMR calculations have been done theoretically and predicted frequency

shifts for the confined superfluid phases [120, 121, 1]. Previous work has been done on

HQVs and NMR signatures in the A-Phase of 3 He [75, 122]. However, it was shown that

the spin-orbit energy destabilizes HQVs in the A-phase, whereas the spin-orbit energy

favors HQVs in the polar phase [123, 124]. Considerable theoretical work has explored

HQVs in the polar phase and non-equal spin pairing phases of confined 3 He [125, 126],

nematic superconductors [127], stability of HQVs in chiral superconductors [128] and

even explored half-quantum vortices in s-wave superconductors [129].

The NMR spectrum of the polar phase of 3 He reveals the signature of half-quantized

topological defects. The order parameter in the polar phase is given by Ap↵i (r) = (r)dˆ↵ m̂i

where the dˆ↵ , m̂i are vectors represents the spin and orbital anisotropy, respectively. The

HQV is stabilized since the spin vector dˆ↵ winds by ⇡ and the orbital vector m̂i winds by ⇡,

thus a total of 2⇡ winding allows the order parameter to be single-valued. Half-quantum

vortices were experimentally detected by cooling 3 He in nematic confined aerogels. The

experiments by Autti et al. [28] stabilizing HQVs in the polar phase of 3 He in Nafen

aerogels were conclusive by observing NMR frequency shifts that detected spin waves

associated with a soliton connecting HQVs. A satellite peak shifted below the bulk
107

Larmor signal was observed when cooling the sample in a rotating cryostat below Tc . The

topological defects were also observed in the presence of no rotation, being explained by

the Kibble-Zurek mechanism, and pinning to the Nafen strands. When the cool-down

below Tc took place, if the applied magnetic field was transverse to the nematic axis, no

soliton frequency shift was observed. However, if the magnetic field was initially zero,

or parallel to the nematic axis and then turned perpendicular below Tc , a satellite signal

was observed in the NMR spectrum indicative of solitons connecting two half-quantized

vortices. The tilting of a magnetic field to generate a soliton between HQVs was first

theoretically predicted in the A-phase [75], however their bound-state spin wave frequency

shifts calculated were incorrect and later corrected [122].

We present here a detailed calculation for HQVs in the confined phases of 3 He in the

presence of a magnetic field. Vortex physics in the presence of a magnetic field is a non-

trivial central research direction both experimentally and theoretically [130, 131, 6, 132,

65, 133, 134]. With recent theoretical advancements on strong-coupling theory [73] and

impurity e↵ects in 3 He, we are able to stabilize the confined superfluid phases and thus

stabilize HQVs in di↵erent phases. We use a square lattice impurity model representing

Nafen-90 line impurities separated by the average interstrand distance [8, 135]. Impurity

e↵ects in vortices play a central role in providing a pinning potential and stabilizing the

superfluid confined phases. The e↵ects of impurities on vortices in p-wave superfluids and

superconductors has been explored in great detail [136, 137, 138, 139, 140, 141].
108

4.1. Ginzburg-Landau Theory

We use the GL theory outlined in the previous chapter to describe HQVs in the

confined phases of superfluid 3 He in Nafen. Here, we modify the previous GL theory by

now adding on an impurity contribution to describe the Nafen impurity lattice [1]. The

impurity contribution to the free energy is given by

XX N
kf2 2
fimp [A] = (~r ~rj ) |A↵i (r)|2 !k zi
48⇡kB Tc ↵ j=1

(4.1) + |A↵i (r)|2 [!? ( xi + zi )]

where the two scattering parameters are given by !k , !? , for scattering along the pz and

in the px py plane, respectively.

To determine equilibrium vortex phases we obtain stationary solutions of the strong-

coupling GL functional, F[A] defined by Eqs. (4.2) - (4.7). The equilibrium state in

a magnetic field are solutions of the Euler- Lagrange equations of F[A] defined by the

functional gradient, G[A] = F/ A† = 0


109

0 = 1 rj rj A↵i + (2 + 3 )ri rj A↵j

↵(T )A↵i 2 1 A⇤↵i T r(AA> ) 2 2 A↵i T r(AA† )

2 3 (AA> A⇤ )↵i 2 4 (AA† A)↵i 2 5 (A⇤ A> A)↵i

XN
kF2 ⇥ ⇤
!k A↵z (~rj ) + !? (A↵x (~rj ) + A↵y (~rj ))
48⇡kB Tc j=1

(4.2) g z H↵ A i H .

Depending on the field orientation, we add on the Zeeman energy term gz H↵ A i H as

shown. Convergence to the steady state solution is strongly dependent on the initializa-

tion of the order parameter since there are many solutions to Eqs. 4.2. We solve Eqs. 4.2

using the L-BFGS numerical optimization algorithm detailed in [142, 65].

4.2. Order Parameter of a Half-Quantum Vortex

We first discuss the order parameter structure for a polar-half quantum vortex. For

an equal-spin pairing state, we choose d~ ? ẑ, and parameterize dˆ as a vector that winds
ˆ = cos ↵(r)x̂ + sin ↵(r)ŷ. Thus the order parameter can be
in the spatial plane only, d(r)

written as

⇣ ⌘
(4.3) (p̂, ~r) = 0e
i#
(i~ y )↵
~
· d(~r) p̂z ,
110

where # = ↵ = ⇡. Thus, for the half-phase quantized vortex # = /2 and the

spin disgyration angle ↵ = /2. Using this parameterization, we can write the order

parameter as
0 1 0 1
i# B
e i↵ 0 C B e
i
0C
(4.4) (p̂, ~r) = 0 e @ A p̂z = 0@ A p̂z .
i↵
0 e 0 1

We can express these results in terms of the 3 He order parameter, A↵i using the decompo-

sition d↵ (~p) = A↵i pi , and demanding A↵i = 0 for i = x, y for the polar phase. We then ob-

tain the spin-up and down components of the gap matrix as "" = ( Axz +iAyz )pz , ## =
i
(Axz + iAyz )pz . For a single HQV, we have "" =| "" (r)|e pz , ## = ## (r)pz . Thus

we now have the spin-up and spin-down order components of a single HQV. Transform-

ing to the order parameter, A↵i , we obtain "" = | (r)|ei pz = ( Axz + iAyz )pz , ## =

| (r)|pz = (Axz + iAyz )pz . We can simplify this to obtain the order parameter profile of a
| (r)| i| (r)|
single HQV as Axz (r) = 2
(1 ei ), Ayz (r) = 2
(1+ei ). Our boundary condition

as r ! 1, in units of p, A ⌘ A/ p, is given by Axz (r ! 1) = 12 (1 ei ), Ayz (r ! 1) =


i
2
(1 + ei ). However, we need to explicitly take into account now that the amplitudes of

the spin-up and down components are not actually the same. Thus a single HQV can be

expressed as a 2 ⇥ 2 gap matrix where the spin-up component is a singly quantized vor-
i
tex, and the spin-down has zero phase winding [118]: "" =| "" (r)|e , ## =| ## (r)|.

Transforming A↵i into the spin-up and down components we find the order parameter of

a single HQV as

1 i i i
(4.5) Axz = ( ## | "" |e ), Ayz = ( ## +| "" |e ),
2 2
111

where the spatial varying OP can be written as | (r)| = tanh(r/⇠). We can generalize

this to a pair of HQVs which is described as a singly quantized vortex in the spin-up

condensate and a singly-quantized vortex in the spin-down condensate. Thus for a pair

of polar HQVs we obtain the more general result

| (r)| ⇥ ⇤
(4.6) A↵i (r) = x̂↵ ẑi ei ei +
iŷ↵ ẑi ei +
+ ei
2

where ± = arctan(y/(x ⌥ x0 )). This will be used as the initialization for a pair of polar

HQVs.

4.3. Numerical Solution of a Single Polar Half-Quantum Vortex

Below we analyze the structure of a single polar HQV in the polar phase of 3 He on

the Nafen-90 impurity lattice. We plot the Cooper pairing density shown in FIG. 4.1 and

it is clear that the HQV is a vortex with a superfluid core. The amplitudes and phases

are shown in FIG. 4.2, clearly indicating that the HQV is a singly-quantized vortex in the

spin-up condensate, and has zero phase winding in the spin-down condensate. To further

understand the vortex structure, we plot the spin up- and down condensates along the

x̂ axis shown in FIG. (5.17) which indicates that the spin-up condensate is zero in the

vortex core as expected for a singly-quantized vortex. Lastly, we plot the magnetization

in FIG. 4.4 which shows the polar HQV is magnetic in its core.
112

800
0.50

0.45
400

0.40
y(nm)

0
0.35

0.30
400

0.25

800
800 400 0 400 800
x(nm)

P
Figure 4.1. Cooper pairing density, ↵,i |A↵i |2 of a single HQV in the polar

phase at p = 15bar, T = 0.95Tc . It is clear the vortex is superfluid in its

core.
113

|C+0 | +0

0.50 ⇡
Amplitude

Phase
0.25 0
|C 0| 0

0 ⇡

Figure 4.2. Amplitude and phase of a single HQV in the polar phase at

T = 0.95Tc , p = 15bar. It can be described a singly-quantized vortex in the

spin-up condensate and a spin-down condensate with zero phase winding.


114

0.5

0.4
|Cµ⌫ |(x, 0)

0.3

0.2

0.1
|C+0 |
|C 0 |
0.0
800 600 400 200 0 200 400 600 800
x(nm)

Figure 4.3. Plots of the polar vortex amplitudes along the x̂ axis in the

polar phase at T = 0.95Tc , p = 15bar. It is clear that the amplitude drops

to zero for the spin-up channel as expected since it hosts a singly-quantized

vortex.
115

0.00 0.0
500

y(nm)
0 0.1

0.05 500 0.2


500 0 500
x(nm)
m(r)/m0

0.10

0.15

0.20
m(x, 0)

800 600 400 200 0 200 400 600 800


r(nm)

Figure 4.4. The spin-polarization along the nematic axis is shown for a

single HQV vortex in the polar phase at T = 0.95Tc , p = 15bar. The

magnetization is maximum in magnitude at the core of the HQV.

4.4. Analytical Solution of a Single Polar Half-Quantum Vortex

To analyze a single polar HQV vortex core and asymptotic structure, we analytically

calculate the solutions of the GL equations neglecting the impurity and Zeeman energy
116

terms. Thus we obtain the equations

0 = 1 rj rj A↵i + (2 + 3 )ri rj A↵j

↵(T )A↵i 2 1 A⇤↵i T r(AA> ) 2 2 A↵i T r(AA† )

(4.7) 2 3 (AA> A⇤ )↵i 2 4 (AA† A)↵i 2 5 (A⇤ A> A)↵i

For a polar HQV, the only non-vanishing components of the OP are Axz , Ayz , thus we

can simplify these equations and write

0 = r2 Axz ↵(T )Axz 2 1 A⇤xz (A2xz + A2yz )

2 2 Axz (|Axz |2 + |Ayz |2 ) 2 3 (A2xz A⇤xz + Axz |Ayz |2 )

(4.8) 2 4 (|Axz |2 Axz + Axz |Ayz |2 ) 2 5 (|Axz |2 Axz + A⇤xz A2yz )

0 = r2 Ayz ↵(T )Ayz 2 1 A⇤yz (A2yz + A2xz )

2 2 Ayz (|Axz |2 + |Ayz |2 ) 2 3 (A2yz A⇤yz + Ayz |Axz |2 )

(4.9) 2 4 (|Axz |2 Ayz + Ayz |Ayz |2 ) 2 5 (|Ayz |2 Ayz + A⇤yz A2xz )

where r2 = @x2 + @y2 . The equations still seem a bit intractable, however we can now
p
make a change of variables by writing Axz = (C+0 + C 0 )/ 2, Ayz = i(C+0 C 0 )/2. The
117

equations we then obtain are considerably easier to handle,

0 = r2 C+0 ↵(T )C+0 4 15 |C 0 |


2
C+0

2
(4.10) 2 234 (|C+0 | + |C 0 |2 )C+0

0 = r2 C 0 ↵(T )C 0 4 15 |C+0 |


2
C 0

2
(4.11) 2 234 (|C 0 | + |C+0 |2 )C 0 .

We can solve these exactly in the vortex core, r ! 0 and the asymptotic limit, r ⇠.

A single half-quantum vortex is described as a singly-quantized vortex in the spin-up

condensate and a spatially varying amplitude in the spin-down condensate. Thus, we can
i
write C+0 = "" (r)e ,C 0 = ## (r). In the vortex core, the spin-up condensate has a

node due to the phase winding, while the spin-down condensate varies spatially with zero

phase winding. The HQV is superfluid since the spin-down channel is non-zero in the

core. Thus in the vortex core, "" (r = 0) = 0 and Eq. 4.10 gives
s
|↵(T )|
(4.12) C 0 (0) = .
2 234

This is just the bulk gap for the -phase. Thus, we have shown that the vortex core is

filled with superfluid -phase and thus given by


0 1
B0 0 1C
1 B C
(4.13) A(r = 0) = p B0 0 iC
2 B C
@ A
0 0 0
118

q
|↵(T )|
where = 2 234
. Since we have calculated the order parameter in the vortex core,

the magnetization at the core of a vortex, r = 0, is polarized along the rotation axis and

given by

|↵(T )|
(4.14) m(r = 0) = m0 ⌦.
2 234

We can now see that the core of a HQV is a spin-polarized ferromagnetic -phase su-

perfluid with intrinsic angular momentum Jzint = +~ arising from the spin state of the

Cooper pairs. This component with zero phase winding can support finite amplitude in

the vortex core which was the basis for the prediction of the ferromagnetic A-core vortex

in the B-phase. The asymptotics of the wavefunction are calculated in a cylindrical basis

as

00 1 0 "" (r) 4 12345 2


(4.15) "" (r) (r) + "" (r) + "" (r) | "" | "" (r) = 0.
r r2 |↵(T )|

Far from the vortex core, r ⇠, the spatial gradients vanish and the spin-up and down

components become equal, we obtain

4 12345
(4.16) "" (r)(1 r 2) | "" (r)|
2
"" (r) = 0.
|↵(T )|

Simplifying we obtain the asymptotic structure of the spin-up condensate given by

p
(4.17) lim "" (r) ⇠ p 1 1/r2
r/⇠ 1

i
p
(4.18) lim C+0 (r) ⇠ pe 1 1/r2 .
r/⇠ 1
119

This is how the spin-up channel, "" (r) behaves in the asymptotic limit r/⇠ 1. We can

write the OP as
0 1
B0 0 1C
p B C
(4.19) A(x ⇠, 0) = p 1 1/x2 B
B0 0 iC
C.
@ A
0 0 0

i
Thus as r ! 1, on the computational boundary, we have just C+0 ⇠ pe as expected.

We now have the asymptotics and the core structure of a half-quantum vortex and sum-

marize that the vortex core is determined to be the superfluid ferromagnetic -phase. We

now proceed to numerical calculations and plot the vortex structure.

The amplitudes of the vortex structure are first plotted along the x̂ axis, shown in

FIG. 4.5, in comparison with our asymptotic analysis given above. We then plot the

amplitudes and phases of a polar HQV shown in FIG. 4.6. As expected, we recover a

singly-quantized vortex in the spin-up channel and a spatially varying zero-phase winding

order parameter in the spin-down channel. Lastly, we plot the transverse currents of the

HQV shown in FIG. 4.7.


120

0.7

0.6

0.5
|Cµ⌫ |(x, 0)

|C+0 |
0.4 |C 0 |
p
p1 1 1/x2
0.3 2

0.2

0.1

0.0
0 100 200 300 400 500 600 700 800
x(nm)

Figure 4.5. It is clear that the asymptotic calculation |C+0 (x ⇠(T ))| ⇠
p
p 1 1/x2 is valid for x 4⇠(T ) ⇡ 400nm.
121

|C+0 | +0

0.707 ⇡
Amplitude

Phase
0.35 0
|C 0| 0

0 ⇡

Figure 4.6. Amplitudes and phases of a single polar half-quantum vortex

at p = 15bar, T = 0.95Tc , on a square grid of ±800nm ⇡ 7⇠(T ). The

core size is ⇡ 30nm. The HQV is a singly-quantized vortex in the spin-

up condensate, and a constant amplitude with zero phase winding in the

spin-down condensate. The core is ferromagnetic superfluid -phase.


122

0.15

0.10

0.05
J(r)/J0

Jy (x, 0)
0.00
Jx (0, y)
0.05

0.10

0.15

800 600 400 200 0 200 400 600 800


r(nm)

Figure 4.7. Transverse mass currents taken at p = 15bar, T = 0.95Tc .

4.5. Pair of Polar Half-Quantum Vortices

We now describe the structure of a HQV pair with a pure polar core in the polar phase

of 3 He confined in Nafen. All calculations reported are on a square grid 1600 ⇥ 1600nm

with grid spacing x = y ⇡ 10nm, impurity spacing r ⇡ 50nm, and in units of the
p
bulk polar phase p = |↵(T )|/2 12345 . We find no noticeable di↵erence in any of the

structure for axial fields ranging from H = 0 ˆ which is expected as a strong


370G ⌦
123

field along the rotation axis favors the dˆ vector to live in the x y plane. However, a

transverse field ranging from H = 0 1500G x̂ destabilizes a HQV pair and collapses into

a singly-quantized vortex for |H| > 1000G. This is expected as the magnetic energy is

minimized for d̂ ? Ĥ, and since we are in an equal-spin pairing phase for d̂ ? ẑ, the HQV

gets suppressed. We first show below the order parameter pairing density and amplitudes.

Below in FIG. 4.8 is a plot of the superfluid Cooper-pairing density for the pair of HQVs

on the impurity lattice.


124

800 0.5

0.4
400

0.3
y(nm)

0
0.2

400
0.1

800 0.0
800 400 0 400 800
x(nm)

P
Figure 4.8. The superfluid density, ↵,i |A↵i |2 = |Axz |2 + |Ayz |2 , is shown

above for the HQV pair at p = 15bar, T = 0.95Tc . The pair of HQVs break

rotational invariance along the x̂ axis and have superfluid, spin-polarized

ferromagnetic cores of size r ⇡ 50nm.

The vortex amplitudes along the x̂ axis below in FIG. 4.9 are plotted in an axi-
P ⇥ ⇤ ⌫
symmetric basis A↵i (r) = p3 µ,⌫ µ↵ Cµ⌫ (r)ei(p µ ⌫) i . The basis is given by quan-

tum numbers µ, ⌫ which are projections of the cooper pair spin and orbital momentum
125

respectively [143]. This plot explicitly shows that a pair of HQVs is given by two singly-

quantized vortices in the spin- up and down condensates.

0.5

0.4
|Cµ⌫ |(x, 0)

0.3

0.2

0.1
|C+0 |
|C 0 |
0.0
800 600 400 200 0 200 400 600 800
x(nm)

Figure 4.9. Amplitude of a polar HQV pair at p = 15bar, T = 0.95Tc along

the x̂ axis. It is clear that a pair of HQVs can be described as a singly-

quantized vortex in both the spin-up and down condensate as both of these

amplitudes vanish in the vortex cores.


126

4.5.1. Magnetic Properties & Mass Currents

The HQVs are superfluid and magnetic in their cores, while the bulk order parameter is

suppressed due to the impurity lattice. The spin polarization is given by

X
(4.20) m(r) = m0 (|C+⌫ |2 |C ⌫ |2 )⌦
⌦ + m?

where m0 = gz0 | P|
2
. For a pair of polar HQVs this simplifies to

(4.21) m(r) = m0 =(Axz A⇤yz A⇤xz Ayz )⌦


as the transverse magnetization, m? , is zero for any ESP phase since d̂ ? ẑ. The mag-

nitude of the magnetization is given by m0 ⇡ n( ~) ln(Ef /kB Tc )( B /Ef )


2
[91, 144].

The magnetization at the core of a vortex, rc , is polarized along the rotation axis and

in the presence of no impurities is given by m(r = rc ) = m0 |↵(T )|/2 234⌦ . The spin-

polarization is generated via the Barnett e↵ect, i.e. rotation induced magnetization. The

intrinsic orbital angular momentum of the Cooper pairs is given by

4m
(4.22) Lorb,k = ✏ijk =(A† A)ij
~

and vanishes for a polar HQV pair since there is no coupling between orbital components

as the orbital anisotropy is purely along ẑi . The magnetic susceptibility of the vortices can

be calculated from the Zeeman energy, ff ield = H↵ ↵ H /2. The change in susceptibility

relative to the normal state is ↵ = 2gz (AA† )↵ and thus ↵ = ↵ N + ↵ =

↵ N 2gz (AA† )↵ . The magnetic susceptibility has already been analyzed in great detail

in the bulk B-phase of 3 He and aerogels [145], however it has not yet been investigated
127

theoretically for HQVs in the polar phase. For a pair of polar HQVs, the magnetic

susceptibility tensor is then given by


0 1
2 ⇤
B1 2g z |A xz | 2g A A
z xz yz 0 C
↵ (r) B C
(4.23) = B 2gz Ayz A⇤xz 1 2gz |Ayz |2 0C
B
C
N @ A
0 0 1

since the only non-vanishing components of the equal-spin pairing phase are Axz , Ayz .

Plots below of the magnetization m(r) and magnetic susceptibility ↵ (r) for the vortices

are given in FIG. 4.10-4.11. The susceptibility is plotted for the components transverse

to the polar ẑ axis, ( xx , xy , yx , yy ); where the o↵-diagonal components have induced

spin polarization transverse to the applied magnetic field. The transverse susceptibilities

are isotropic as the HQV is a ferromagnetic phase, and the largest principal value of the

susceptibility is zz which corresponds to the magnetic field direction.

By carrying out a boost transformation on the gradient terms of the free energy, we

obtain the superfluid mass current density in the rest frame of the excitations,

(4.24) ji = j0 = A⇤↵j rj A↵i + A⇤↵j ri A↵j + A⇤↵i rj A↵j .

The transverse anisotropic mass currents are in units of j0 = 4m/~ and shown in FIG.

4.12. Unlike the polar-distorted chiral HQV, the ẑ-axis current density vanishes for a pure

polar HQV, jz = 0.
128

0.2
0.2 500 0.1

y(nm)
0 0.0
500 0.1
0.2
0.1
500 0 500
x(nm)
m(r)/m0

0.0

0.1

0.2
m(x, 0)

800 600 400 200 0 200 400 600 800


r(nm)

Figure 4.10. Vortex magnetization profile along the ẑ axis m(r) for the

HQVs at p = 15bar, T = 0.95Tc . The spin-polarization is maximum in

magnitude in the vortex cores. Inset: Density plot of the same.


129

1.00

0.95
N

0.90
(r)/

0.85

0.80 xx (x, 0)

xx (0, y)
0.75 yy (x, 0)

yy (0, y)
0.70
800 600 400 200 0 200 400 600 800
r(nm)

Figure 4.11. Diagonal components of the zero-field transverse susceptibility

for the HQVs at p = 15bar, T = 0.95Tc in units of the normal-state N. As

shown in Eq. 4.23 , we have zz = N, thus we need to only consider the

components in plane. As we move away from the vortex cores along the x̂

axis, the vortices are more easily polarized which is expected as superfluidity

gets stronger as we move away. In the vortex cores, the vortices are equally

polarizable along either field direction and thus xx (x, 0) ⇡ yy (x, 0). The

o↵-diagonal are negligible compared with the diagonal components and sat-

isfy xy = yx .
130

100

50
y(nm)

50

100
600 400 200 0 200 400 600
x(nm)

Figure 4.12. Expanded view of the mass currents taken at p = 15bar, T =

0.95Tc clearly showing the broken axial symmetry of the vortex. It can

be seen that the vortex structure near the half-quantum vortex cores is

the source of the anisotropic current density. The currents vanish near the

zero-phase winding polar cores. Currents are scaled in units of j0 defined

by Eqn. 13.

4.6. Conversion of Half-Quantum Vortices into Singly-Quantized Vortices

We now investigate the energetics and self-induced Barnett magnetization of HQVs

and singly-quantized vortices (SQVs). We first calculate the magnetization as a function

of the intervortex spacing which is shown in FIG. 4.13. The intervortex spacing in the

equilibrium solutions is dependent on the separation in the initial condition. It is clear

that the Barnett e↵ect decreases as the separation gets smaller and eventually vanishes for

an SQV. This is expected as the spin-up and down components are identical for an SQV

and thus there is no net magnetization. We find that the Barnett induced magnetization

of the HQVs in the polar phase are 4-5 orders of magnitude smaller than the normal-core
131

vortex in the Bulk B-phase. Despite the normal-core also being zero in the vortex core

as seen in individual spin condensates of HQVs, the magnetization is much larger in the

normal-core vortex due to the vast di↵erence in structure of the vortex order parameter,

A↵i . We also plot the amplitudes and phases of the SQV arising from the collapse of a

HQV pair shown in FIG. 4.14. We then analyze the energetics of a pair of HQVs as a

function of inter-vortex spacing which is shown in FIG. 4.15. We determine that the polar

HQV is indeed a local minima of the GL functional, and not a metastable phase. The

SQVs are the highest energy vortex phase that we stabilize in the polar phase.
132

0.2 0.2
500 0.1

y(nm)
0 0.0
500 0.1
0.1 0.2
500 0 500
x(nm)
m(r)/m0

0.0
mxi =±500nm (x, 0)
mxi =±450nm (x, 0)
mxi =±400nm (x, 0)
mxi =±350nm (x, 0)

0.1 mxi =±300nm (x, 0)


mxi =±250nm (x, 0)
mxi =±200nm (x, 0)
mxi =±150nm (x, 0)
mxi =±100nm (x, 0)

0.2 mxi =±50nm (x, 0)


mSQV (x, 0)

800 600 400 200 0 200 400 600 800


r(nm)

Figure 4.13. Spontaneous zero-field magnetization as a function of the in-

tervortex spacing set by the initial condition of a pure polar HQV. The

magnetization is generated by the Barnett e↵ect, i.e. rotation induced. It

is clear that the Barnett e↵ect vanishes for a singly-quantized vortex. Note,

the Barnett induced magnetization of the vortices in the polar phase are

4-5 orders of magnitude smaller than the normal-core vortex in the Bulk

B-phase as described in detail in the text.


133

|C+0 | +0

0.50 ⇡
Amplitude

Phase
0.25 0
|C 0| 0

0 ⇡

Figure 4.14. The collapse of a HQV pair into a conventional singly-

quantized vortex. Both spin-up and down components drop to zero in the

core and have 2⇡ phase winding as expected.


134

2.066

2.064
F/F0

2.062

2.060

2.058

2.056
0 200 400 600 800 1000
x(nm)

Figure 4.15. It is clear that as the intervortex separation x increases, the

HQV pair energy is lowered. A singly-quantized vortex, x = 0, is higher

in energy than a pair of HQVs and thus the HQV is not a metastable phase,

but a local minimum of the free energy. For x > 1000nm, the energy starts

to become larger as the stable phase is now the B phase and no longer the

confined polar phase.


135

4.7. Magnetic Field Evolution of a Polar Half-Quantum Vortex

Lastly, we analyze the polar HQVs as a function of a magnetic field. There is no

discernible di↵erence in the vortex structure in axial fields at all for fields ranging Ĥ =

0 500G ⌦. However, when analyzing the HQVs in a transverse magnetic field we find

a very di↵erent result. We plot the evolution of the polar component, <[Axz ](x, 0), along

the vortex core axis x̂ in a transverse magnetic field, Ĥ = 0 1500G x̂, shown in FIG. 4.16.

As expected, the polar component with d̂ = x̂ is suppressed since the magnetic energy is

minimized for d̂ ? Ĥ, thus Ĥ = d̂ = x̂ is an unfavorable energy state. The inter-vortex

separation approaches zero and thus the pair of HQVs collapses into a singly-quantized

vortex as the transverse field is increased.


136

H = 0G
0.6 H = 100G
H = 200G
H = 300G
0.5 H = 400G
Re[Axz ](x, 0)

H = 500G
H = 600G
0.4 H = 700G
H = 800G
H = 900G
0.3 H = 1000G
H = 1100G
H = 1200G
0.2 H = 1300G
H = 1400G
H = 1500G
0.1

0.0
0 100 200 300 400 500 600 700 800
x(nm)

Figure 4.16. Evolution of the polar component, <[Axz ](x, 0), along the vor-

tex core axis x̂ in a transverse magnetic field, Ĥ = x̂. The magnetic energy

is minimized for d̂ ? Ĥ, hence the suppression of the order parameter. For

H > 1000G, the HQV pair approaches that of a singly-quantized vortex

centered near x = 0 as shown.


137

4.8. Chiral Phase Order Parameter

We now discuss the confined polar-distorted chiral order-parameter and the supercur-

rents generated by the impurity lattice arising from the intrinsic coupling between di↵erent

orbital angular momentum directions of the Cooper pairs. Thus the supercurrents in the

confined phases are generated spontaneously along the rotation axis, despite having zero

phase gradient and no mass transport along the rotation axis. This only occurs for the

polar-distorted chiral phase as the pure polar phase has an orbital lobe along p̂z only.

Thus, we see that the manifestation of the spontaneous supercurrents in the HQVs are

even observed in the polar-distorted chiral phase. We calculate the order parameter and

supercurrents on an impurity lattice with di↵erent order parameters given by pz + ipx and

pz + ipx + ipy . For the orbital component of the gap given by pz + ipx , we show our results

in FIG. 4.17. Our calculations for pz + ipx + ipy are shown in FIG. 4.18.
138

|Azx |/ c |Azz |/ c

0.305 0.795
0.300 0.790
|Azx | Amplitude

|Azz | Amplitude
0.295 0.785
0.780
0.290
0.775
0.285
|Azy |/ c jz /j0 0.770
0.280
0.765
0.275 0.760
0.270 0.755

Figure 4.17. Order parameter amplitude and spontaneous supercurrents

generated by the polar-distorted chiral phase confined in Nafen. Here, we

initialize a chiral phase of the form pz + ipx . The spontaneous supercurrents

are generated by the coupling of p̂x , p̂z and integrate to zero as expected.
139

|Azx |/ c |Azz |/ c

0.305 0.795
0.790
|Azix+iy | Amplitude

0.300

|Azz | Amplitude
0.295 0.785
0.780
0.290
0.775
0.285
|Azy |/ c jz /j0 0.770
0.280
0.765
0.275 0.760
0.270 0.755

Figure 4.18. Order parameter amplitude and spontaneous supercurrents

generated by the polar-distorted chiral phase confined in Nafen. Here, we

initialize a chiral phase of the form pz + i(px + py ). The spontaneous super-

currents are generated by the coupling of p̂x , p̂y , p̂z and again integrate to

zero as expected.
140

4.9. Pair of Polar-distorted Chiral Half-Quantum Vortices

We now cool in temperature into the polar-distorted chiral A phase and discover a new

HQV pair. The structure of a HQV pair with a polar-distorted chiral core is discussed

below. We find that this HQV pair nucleates a chiral phase in its core in addition to the

polar phase. The chiral phase observed is stable in the x̂ ŷ plane and breaks rotational

symmetry. A similar phase in confined 3 He was observed in 100nm cylindrical pores [73].

We initialize this chiral vortex by maintaining the dominant pz orbital, but introducing

a sub-dominant orbital component along px . Thus we replace the polar HQV pair initial

condition by writing ẑi 7! ẑi + i✏x̂i , where the chiral axis is along ŷi and ✏ controls the

chiral distortion. We obtain the new initial condition as

(r) ⇥ ⇤
A↵i (r) = x̂↵ ẑi ei ei + iŷ↵ ẑi ei + ei +

2
(r) ⇥ ⇤
(4.25) +✏ ŷ↵ x̂i ei + ei + + ix̂↵ x̂i ei ei +
.
2

The boundary conditions are then imposed on these initial conditions with the phase

windings shown. Note, the phase ± is the phase for the location of the two di↵erent

vortices. If we distort the orbital components of the order parameter further, using

ẑi 7! ẑi + i✏x̂i + i✏ŷi , we obtain

(r) ⇥ ⇤
A↵i (r) = x̂↵ ẑi ei ei + iŷ↵ ẑi ei + ei +

2
(r) ⇥ ⇤
+✏ ŷ↵ x̂i ei + ei + + ix̂↵ x̂i ei ei +
+
2
(r) ⇥ ⇤
(4.26) +✏ ŷ↵ ŷi ei + ei + + ix̂↵ ŷi ei ei +
+
2
141

In the limit ✏ = 0, we recover the pure polar half-quantum vortex pair as expected. The

phase of each singly quantized vortex is given by ± = arctan(y/(x ⌥ x0 )).

Both initializations of the OP were used, however the converged equilibrium solution

is always a HQV pair with orbital angular momentum along px , py , pz . Thus using the

initialization in Eq. (4.26) improves convergence speed greatly. Thus, the pair of HQVs

break rotational symmetry about the x̂ axis and has superfluid cores hosting a chiral and

polar phase. The situation of a polar-distorted chiral vortex core is similar to the chiral

phase of a 100nm pore in 3 He with zero phase winding. There are two possible chiral

phases that can nucleate from the polar phase, we only find a vortex phase in which

the chiral axis lives in the xy plane and breaks rotational symmetry. There is also the

possibility of a chiral axis in the radial direction with a pure polar core, but we do not

find this to be an equilibrium solution. Thus the chiral vortex in the x̂ ŷ plane has

dominant chiral core components given by Axx , Ayx , Ayy . Note, the chiral and polar phase

bulk OP are given by


s s s
|↵|(T ) |↵|(T ) ˜ cp = p 345
(4.27) p = , c = , = ,
2 12345 2 345 c 12345

at T = 0.85Tc , p = 15bar, ˜ cp ⇡ 0.891.

4.9.1. Intrinsic Vortex Orbital and Spin Angular Momentum

We now consider a polar-distorted chiral HQV pair using the initialization from Eq. (4.26)

in the limit ✏ ⌧ 1. All plots below are shown for ✏ = 0.25, H = 0G, T = 0.85Tc , p = 15bar,
p
in units of the bulk chiral phase c = |↵(T )|/ 345 and with the same impurity and grid

spacing described earlier. The converged value of ✏ is independent of the initial value
142

set, as long as ✏ ⌧ 1. The spin-polarization along the rotation axis is shown FIG.

4.19, the transverse components of the magnetization density are zero similarly to the

polar HQV pair as this vortex pair is also an ESP phase. We also calculate the intrinsic

orbital angular momentum,Lorb,k . The intrinsic orbital angular momentum of the Cooper
4m
pairs is given by Lorb,k = ~
✏ijk =(A† A)ij and is non-vanishing for a polar-distorted chiral

HQV since there is now coupling between the di↵erent orbital degrees of freedom. Thus,

unlike the polar HQV pair, we observe that the intrinsic orbital angular momentum for

a polar-distorted chiral HQV pair is non-zero and is shown for all three components, Lk ,

in Fig. 4.20. The orbital angular momentum component, Lz specifically shows the chiral

nature of the vortex cores.


143

0.4 0.4
500 0.2

y(nm)
0.3 0 0.0
500 0.2
0.2 0.4
500 0 500
x(nm)
m(r)/m0

0.1

0.0

0.1

0.2

0.3
m(x, 0)
0.4
800 600 400 200 0 200 400 600 800
r(nm)

Figure 4.19. Vortex magnetization along the ẑ axis m(r) for the HQVs at

p = 15bar, T = 0.85Tc as the transverse magnetization is zero since it

couples to d̂ = ẑ, but d̂ ? ẑ. The magnitude of the spin-polarization is

maximal in the cores. It is clear that the vortex cores are superfluid and

magnetic. Inset: Density plot of the same.


144

1.00 0.150
500

y(nm)
0.75 0 0.075
500
0.50 0.000
500 0 500
x(nm)
Lk (r)/L0

0.25

0.00

0.25

0.50
Lx (0, y)
0.75 Ly (x, 0)
Lz (x, 0)
1.00

800 600 400 200 0 200 400 600 800


r(nm)

Figure 4.20. Intrinsic orbital angular momentum, Lk , for the HQVs at p =

15bar, T = 0.85Tc . We note that Lk is on the order of Lorb (n~/4)( /Ef )2 ,

with orb ⇠ O(ln(Ef /kB Tc )). Inset: Density plot of Lz projecting out

the components with orbital angular momentum x̂i , ŷi indicating the chiral

nature of the vortex core Cooper pairs.


145

4.9.2. Chiral Vortex Mass Currents

We now discuss the transverse currents in the polar-distorted chiral HQV. We first plot

the transverse mass currents which are shown in units of j0 = 4m/~ in FIG. 4.21. Unlike

the polar core HQV, there is a spontaneous axial supercurrent along the polar, ẑ, axis for

the polar-distorted chiral HQVs. This axial supercurrent is embedded in the HQV in the

London limit as it purely arises from the coupling between di↵erent orbital directions of

the Cooper pairs. The axial supercurrent is given by js,z = j0 =(A⇤↵j rj A↵z + A⇤↵z rj A↵j )

and was discussed in detail in Chapter 4 of this thesis.

100

50
y(nm)

50

100
400 200 0 200 400
x(nm)

Figure 4.21. Anisotropic mass currents taken at p = 15bar, T = 0.85Tc .

The transverse mass currents are given by j(r) =

j0 = [A⇤↵z rx A↵z x̂ + A⇤↵z ry A↵z ŷ] .

4.10. Superfluid Vortex Phase Diagram of 3 He Confined in Nafen

The confined superfluid phases of 3 He are stabilized by infusing 3 He into Nafen-90

nematic line impurities. We model the impurities by using a Nafen-90 impurity lattice

model. We calculate the vortex phases within the confined phases on the impurity lattice
146

and summarize our results in FIG. 4.22. We start by calculating the polar phase order

parameter on the impurity lattice as shown in FIG. 4.23 and obtain excellent agreement

with previous calculations for |H| = 0G [1]. For H = 100 ˆ , we compare our
370G ⌦

results with the experimentally reported phase transitions shown in FIG. 4.10. In order

to compare with the experimental data, we analytically calculate the normal to polar

phase transition temperature as a function of pressure, and numerically calculate the

phase boundary separating the polar-distorted chiral phase and polar phase. We note

that our theoretical strong-coupling model accurately accounts for the relative stability

of the confined phases in agreement with experimental data; which di↵ers from the re-

cent calculation done in weak-coupling [126] which is in strong disagreement with the

experimentally observed phase transitions. The polar phase order parameter is given by

Ap↵i (r) = (r)ẑ↵ ẑi , while the polar-distorted chiral phase order parameter is given by

ACh
↵i (r) = (r)ẑ↵ ẑi + i Ch (r)ẑ↵ x̂i . The transition temperature for the normal to polar

phase is obtained analytically by linearizing Eqs. 4.2 in the polar limit, A↵i = Azz ẑ↵ ẑi ,

we find that

T ⇡ 2 ⇠0 (p)!k
(4.28) =1
Tc 4L2

where L is the interstrand aerogel distance, and ↵0 = Nf /3.


147

30 Chiral-Polar, H=0G, Theory


Chiral-Polar, H=300G, Theory
Normal-Polar, H=0G, Theory
25 Normal-Polar, H=300G, Theory
Chiral-Polar, H=100G-370G, Experiment
Normal-Polar, H=100G-370G, Experiment

20 |A i |2 Polar-Distorted
Polar Phase
p (bar)

0.8 Chiral A-Phase


500
0.7
15
y(nm)

0 0.6
0.5
500 0.4 |A i |2
10 500 0
x(nm)
500
0.3
500
y(nm)
0.5

0 0.25
5 500
Normal
0 Phase
500 0 500
x(nm)
0
0.7 0.8 0.9 1.0
T /Tc

Figure 4.22. Vortex phase diagram for the equal-spin pairing phases of con-

fined superfluid 3 He. We calculate the structure of a polar HQV pair in the

polar phase at p = 15bar, T = 0.95Tc and then cool in temperature into the

polar-distorted chiral A phase to stabilize a pair of polar-distorted chiral

HQVs at p = 15bar, T = 0.85Tc . The vortex phases are calculated in equi-

librium and compared with the experimental equilibrium phase transitions

[7]. Insets: Polar-distorted chiral HQV pair and Polar HQV pair Cooper

pairing densities.
148

| zz |/ p
100
0.72

50
0.71
y(nm)

0
0.70

50 0.69

100 0.68
100 50 0 50 100
x(nm)

Figure 4.23. Polar phase order parameter suppressed by a lattice of square


p
impurities in Nafen-90 in units of bulk P = |↵|/2 12345 at T =

0.95Tc , p = 15 bar. The impurities for Nafen-90 have interstrand distance

of r ⇡ 50nm , and the scattering cross-section eigenvalues are given by

!? = 4.25nm, !k = 1.073nm [8].

Note, the confined superfluid 3 He phase diagram is shown in an axial magnetic field.

All phase transitions shown were calculated in thermodynamic equilibrium. We compare

our results with experimentally reported phase transitions in a magnetic field of H =


149

100 ˆ [7]. Note, the transition lines in a magnetic field are shifted to the left of
370G ⌦

the zero-field line since the configuration Ĥ = d̂ = ẑ is energetically unfavorable as the

magnetic energy is minimized for Ĥ ? d̂. Having confirmed that we are in the correct

phases, the vortex structures are now calculated on the Nafen-90 impurity lattice shown

in FIG. 4.23. The two pairs of HQVs we find are a pure polar HQV, and a polar-distorted

chiral HQV pair. Initialization and boundary conditions for the polar-distorted chiral

HQV pair are slightly modified as described later. A polar core HQV pair is first calculated

and explained below. Plots below show the order parameter and current structure of the

HQV pairs.
150

|C 0 | |C+0 |
0.50

0.25

0
0 +0


100

50
y(nm)

50

100
600 400 200 0 200 400 600
x(nm)

Figure 4.24. Amplitudes, phases and transverse mass currents of a polar

HQV pair at p = 15bar, T = 0.95Tc , on a square grid of ±800nm ⇡ 7⇠(T ).

A pair of HQVs can be described as a singly-quantized vortex in both

the spin- up and down condensate, thus these are the only two non-zero

components of the order parameter as shown. The suppression of the order

parameter outside the vortex cores is from the Nafen-90 impurities.


151

|C++ | ++ |C+ | + |C+0 | +0


0.6
Amplitude

0.3

0

|C +| + |C | |C 0 | 0
Phase

100
y(nm)

100
600 400 200 0 200 400 600
x(nm)

Figure 4.25. Amplitudes, phases and transverse mass currents of a polar-

distorted chiral HQV pair at p = 15bar, T = 0.85Tc , on a square grid of

±800nm ⇡ 13⇠(T ). It is clear that this pair of HQVs nucleate a chiral

phase in its core as shown in the C++ , C + components. The dominant

bulk components are the polar components, C+0 , C 0 , as expected. The

components with d̂ k ẑ are zero since the HQVs are an equal-spin pairing

phase. The vortex cores are less circular than the pure polar HQV pair and

more elliptical which is a manifestation of the underlying chiral phase order

parameter on the impurity lattice.


152

We also plot the axial supercurrent along the ẑ axis as shown below. We now discuss

the transverse currents and axial supercurrent in the polar-distorted chiral HQV. We first

plot the transverse mass currents which are shown in units of j0 = 4m/~. Unlike the

polar core HQV, there is a spontaneous axial supercurrent along the polar, ẑ, axis for

the polar-distorted chiral HQVs. This axial supercurrent is embedded in the HQV in the

London limit as it purely arises from the coupling between di↵erent orbital directions of

the Cooper pairs. The axial supercurrent is given by js,z = j0 =(A⇤↵j rj A↵z + A⇤↵z rj A↵j )

and shown below.


153

800

600 0.04

400
0.02
200
y(nm)

0 0.00

200
0.02
400

600 0.04
800
800 600 400 200 0 200 400 600 800
x(nm)

Figure 4.26. Spontaneous supercurrent flowing along the polar axis ẑ, in

units of j0 , despite having zero phase gradient along this axis. The super-

current integrates to zero as expected as there is no mass transport along

this axis. These currents were calculated at p = 15bar, T = 0.85T c.

The HQVs are superfluid and magnetic in their cores, while the bulk order parameter

is suppressed due to the impurity lattice. The spin polarization is given by m(r) =
P
⌦ + m? where m0 = gz0 | P |2 . For a pair of polar HQVs this
m0 ⌫ (|C+⌫ |2 |C ⌫ |2 )⌦

simplifies to m(r) = m0 =(Axz A⇤yz A⇤xz Ayz )⌦


⌦ as the transverse magnetization, m? , is
154

zero for any ESP phase since d̂ ? ẑ. The magnitude of the magnetization is given by

m0 ⇡ n( ~) ln(Ef /kB Tc )( B /Ef )


2
[91, 144]. The spin-polarization is generated via the

Barnett e↵ect, i.e. rotation induced magnetization. We plot these results for the HQVs

below.
155

0.3
0.2 mC (x, 0)
500
mp (x, 0)

y(nm)
0 0.0
0.2
500 0.2
500 0 500
0.1 x(nm)
m(r)/m0

0.0

0.1 0.2
500
y(nm)

0 0.0
0.2 500
0.2
500 0 500
x(nm)
0.3

800 600 400 200 0 200 400 600 800


r(nm)

Figure 4.27. Vortex magnetization profile along the ẑ axis m(r) for the

HQVs. The polar HQV pair is taken at p = 15bar, T = 0.95T c, while

the chiral pair of HQVs is taken at p = 15bar, T = 0.85Tc . The spin-

polarization is maximum in magnitude in the vortex cores. The transverse

magnetizations are zero since they couple to dˆ = ẑ, but dˆ ? ẑ. It is clear

that the vortex cores are superfluid and magnetic. Inset: Density plots of

the same.
156

4.11. Conclusions

Half-quantum vortices in the equal-spin pairing confined phases of 3 He were analyzed

by calculating steady state solutions to the strong-coupling Ginzburg Landau equations in

the presence of nematic aerogel line impurities and a magnetic field. We determined the

regions of stability for equilibrium HQV phases in the confined phases of 3 He in Nafen;

a polar HQV pair and a polar-distorted chiral HQV pair. The structure of a polar HQV

pair was investigated in the polar phase of 3 He. A polar-distorted chiral core was stabi-

lized upon cooling from the polar phase into the polar-distorted chiral A phase. We also

calculated the confined superfluid phase diagram in a magnetic field and obtain excellent

agreement with the experimentally reported phase transitions. All calculations have been

performed as a function of both axial and transverse magnetic fields. We calculated the

energetics and magnetic properties of each vortex and compared with the crossover to

singly-quantized vortices. The Barnett-induced magnetization is carefully analyzed as a

function of inter-vortex spacing. We conclude with a few forward looking observations

to further understand the HQV structures. Further development in understanding the

soliton NMR signatures binding the HQVs can be based on using a random impurity

model which can further provide a stronger pinning potential for the HQVs as needed for

NMR comparisons [1]. We can then consider time-dependent Ginzburg-Landau theory as

dynamics are needed to obtain NMR signatures under non-equilibrium conditions. There

has been microscopic theories to understand the spin dynamics of the NMR signatures

for the bulk superfluid phases of [51, 52, 146] and recently a development for the NMR

signatures in the polar phase [147]. By considering the dynamics of a pair of HQVs in an

axial field and then tilting the field to the transverse direction, we can generate a soliton
157

under non-equilibrium conditions which acts as the glue binding a pair of HQVs. The

NMR signatures of a soliton can then be calculated and compared with the experimental

reported frequency shifts [28, 148]. In general, the non-equilibrium analysis of moving

vortices [149, 150, 151, 152, 153] and dynamical vortex phases transitions is a highly

complicated direction [154, 155, 133, 156, 157, 158]. Secondly, impurity e↵ects play

a strong role on bound states within vortex cores [159, 160] which can be further inves-

tigated as HQV cores are known to host Majorana fermions. It is clear from our work

that developing an impurity model is the correct way to understand the confined phases

in 3 He. Lastly, while vortex charging mechanisms are well understood in superconductors

and the bulk phases of 3 He [161, 162, 139, 140, 163, 164, 165, 141], they have not

been studied in the confined phases of superfluid 3 He. There are a multitude of charging

e↵ects that can a↵ect the underlying structure of the vortex cores [166, 167, 168], thus

studying charging e↵ects on HQVs is an interesting phenomena that extends outside the

realm of strong-coupling GL theory.


158

CHAPTER 5

Quasiparticle Bound States in Polar Half-Quantum Vortices

We develop an impurity based model to analyze the quasiparticle bound state spectrum

of half-quantum vortices (HQVs) in the confined phases of superfluid 3 He in the strong-

coupling limit. We consider weak-coupling p-wave superfluid 3 He in equilibrium. Half-

quantum vortices have not yet been studied microscopically in 3 He, however in other

superconducting systems fractional vortices have been proposed to exist and investigated

theoretically [169, 170]. HQVs have been studied using the quasiclassical theory recently

in s-wave Niobium bi-layer superconductors [129]. The quasiclassical equations of motion

are given by Eilenberger’s equation as described in detail in Chapter 2. Before solving

the QC equations numerically, we perform the analytical continuation in order to write

the Riccati equations in terms of the retarded Green’s functions, i✏n 7! ✏ + i0+ . We

consider the Matsubara energies satisfying [ ✏/ , ✏c / ] discretized in units of 0.1 . The

cuto↵ is determined at each temperature in terms of a fixed cuto↵ energy, ✏c ⇠ 40kB Tc ,

✏n = (2n + 1)⇡kB T . Thus we need to solve this equation with an input = GL , set

the cuto↵ energy, and then solve the ordinary di↵erential equation using Runge-Kutta

techniques [171, 172].

5.1. Local Density of States

We solve equations Eq. (2.15) assuming a polar phase gap structure for the mean-field

order parameter as the initial condition. Recall, the order parameter can be cast into the
159

form A↵i = dˆ↵ ẑi . Thus for dˆ ? ẑ we can write the mean-field gap matrix as
0 1
B "" 0 C
(5.1) ↵ (~pF , ~r) = @ A,
0 ##

where "" = ( Axz + iAyz )pz , ## = (Axz + iAyz )pz . Thus, using our GL solutions for

the equilibrium polar HQV solution, A↵i , we calculate the gap structure and solve the

Riccati equations to obtain the Riccati amplitude and Green’s functions using Eqs. (2.38),

(2.39). Having calculated the Green’s functions, we use the gap equation to obtain the

updated self-consistent order parameter, . In the self-consistency equation, we assume

the orbital structure of the gap is three dimensional having all non-zero components of

px , py , pz . We now calculate and plot the local fermionic spectral density of states (LDOS)

given by Eq. (2.27), respectively. The fermionic local density of states around vortex

structures exhibiting unconventional pairing has been studied numerically and analytically

in great detail [173, 174, 175]. Numerical implementation methods to greatly speed up

convergence of the self-consistency equations has also been discussed as of recent [176].

We now show results for the density of states in the figures below.
160

0.0175

0.0150
N (x, 0, p̂ = 0, ✏ = 0)
0.8
0.0125 500
0.7

y(nm)
0 0.6
0.0100
0.5
500
0.4
0.0075
500 0 500
x(nm)
0.0050

0.0025

0.0000

800 600 400 200 0 200 400 600 800


r(nm)

Figure 5.1. Zero-energy local density of states peaked in the vortex cores

along the x̂ axis. Plot taken at p = 15bar, T = 0.95Tc . Inset: Cooper

pairing density for a pair of HQVs.

5.2. Spin-Resolved Density of States

We now calculate and plot the spin-resolved density of states, using Eq. (2.31, for a

pair of polar half-quantum vortices.


161

0.0175

0.0150
N ""(x, 0, p̂ = 0, ✏ = 0)
0.8
0.0125 500
0.7

y(nm)
0 0.6
0.0100
0.5
500
0.4
0.0075
500 0 500
x(nm)
0.0050

0.0025

0.0000

800 600 400 200 0 200 400 600 800


r(nm)

Figure 5.2. Spin-resolved zero-energy local density of states peaked in the

spin-up condensate along the x̂ axis. Plot taken at p = 15bar, T = 0.95Tc .


162

0.0175

0.0150
N ##(x, 0, p̂ = 0, ✏ = 0)
0.8
0.0125 500
0.7

y(nm)
0 0.6
0.0100
0.5
500
0.4
0.0075
500 0 500
x(nm)
0.0050

0.0025

0.0000

800 600 400 200 0 200 400 600 800


r(nm)

Figure 5.3. Spin-resolved zero-energy local density of states peaked in the

spin-down condensate along the x̂ axis. Plot taken at p = 15bar, T =

0.95Tc .

5.3. Conclusion

Using the quasiclassical Eilenberger equation, I presented self-consistent calculations

of the structure and excitation spectra of half-quantum vortices in an impurity model

of the superfluid phases of 3 He in Nafen. The quasiclassical calculations are all done in

weak-coupling, thermodynamic equilibrium and zero magnetic field. In addition to the

local density of states for a pair of polar-phase half-quantum vortices I report results for

the spin-resolved density of states showing that the fermionic excitations are confined to

either the spin-up or spin-down condensates depending on whether the HQV corresponds

to the "" or ##, respectively.


163

CHAPTER 6

Conclusion

We developed a theoretical model to construct a phase diagram that accurately de-

scribes an experimentally observed phase transition of vortices in the B-phase of rotating

superfluid 3 He. Using a strong-coupling Ginzburg-Landau (GL) theory, we calculated a

phase diagram to account for the relative stability between di↵erent vortex states. We

determined the pressure-temperature phase diagram in the presence of a magnetic field to

classify the regions of stability for the double-core and A-phase-core vortices in rotating
3
He-B. We calculated the metastability curve, the equilibrium transition line and com-

pared our results with experimental data. The structure of the vortices is also computed

and compared with previous theoretical work. Numerical methods used to efficiently

solve the GL equations are outlined as well. Our theory is able to predict a vortex phase

transition far away from T ⇡ Tc (p), which was previously the only region of validity for

the GL theory. We demonstrate the large computability power within our model as our

results agree well with experimental data. This is an indication that the strong-coupling

GL theory can now be used to describe a vast amount of phenomena in low temperature

physics.

After investigating singly-quantized vortices in the B-phase, we studied half-quantum

vortices in the superfluid phases of 3 He confined in Nafen. We calculated the phase di-

agram for the di↵erent confined phases by accurately determining second-order phase

transitions, and calculated the vortex structures within each phase. Again, based on a
164

theoretical model that accurately accounts for the experimentally observed confined su-

perfluid phases of 3 He, we developed a theoretical model to describe the formation of

half-quantum vortices (HQVs) within these confined phases. HQVs form exotic topo-

logical phases of matter that have only recently been discovered experimentally. Our

theoretical phase diagram is in excellent agreement with recent experimental data, and

within these phases we furthermore discover two di↵erent pairs of exotic HQVs. A pair of

polar HQVs was discovered in the polar phase in agreement with the recent experiment

describing the formation of HQVs in the polar phase. We also predict a newly discovered

half-quantum vortex in a phase of matter that has been recently realized experimentally.

Our phase diagram and vortex phases are signatures of new phases of matter, and also

indicate the strength of our theory to provide such accurate experimental agreement. It

is clear from our theory on vortices and phase transitions that it can be widely applicable

in studying a broad range of inhomogeneous phases away from T ⇡ Tc (p). Furthermore,

it is clear from our work that we can describe exotic half-quantum vortices providing

direct signatures in agreement with recent experiments. Thus, the theoretical models we

developed are able to explain experimental phenomena and predict new phases of matter.

Our model is the first to calculate the bulk and vortex phase diagram in multiple phases

of superfluid 3 He as a function of temperature, pressure and magnetic field. Thus, we

calculated the structure of many di↵erent vortices in the superfluid phases of 3 He, and

discovered new phases of matter confined in the mesoscopic cores of quantized vortices.

We lastly calculated the microscopic structure of half-quantized vortices in the context

of weak-coupling Eilenberger theory to understand the bound-state spectrum of HQVs.


165

The bound state spectrum is supported by the current flow around HQVs and the quasi-

particle local density of states. The spin-resolved density of states is also calculated in

each vortex core indicating singly-quantized vortices in di↵erent spin sectors. For future

outlook, we can develop a strong-coupling model in order to stabilize the polar-distorted

chiral A phase and thus investigate the structure of polar-distorted chiral HQVs.
166

References

[1] J. J. Wiman. Ph.D Thesis : Quantitative Superfluid Helium-3 From Confinement

to Bulk. 2019.

[2] J. C. Wheatley. Experimental Properties of Superfluid 3 He. Rev. Mod. Phys., 47:415,

1975.

[3] Dennis S Greywall. 3 He specific heat and thermometry at millikelvin temperatures.

Phys. Rev. B, 33:7520–7538, 1986.

[4] V. Goudon. Magnétisme nucléaire de l’3He liquide: nouvelle détermination du

paramètre de Landau F0a . PhD thesis, Université Joseph-Fourier - Grenoble I, 2006.

[5] Pekola, J. P. and Simola, J. T. and Hakonen, P. J. and Krusius, M. and Lounasmaa,

O. V. and Nummila, K. K. and Mamniashvili, G. and Packard, R. E. and Volovik, G.

E. Phase Diagram of the First-Order Vortex-Core Transition in Superfluid 3 He-B.

Phys. Rev. Lett., 53:584–587, Aug 1984.

[6] Kasamatsu, Kenichi and Mizuno, Ryota and Ohmi, Tetsuo and Nakahara, Mikio.

E↵ects of a magnetic field on vortex states in superfluid 3 He B. Phys. Rev. B,

99:104513, Mar 2019.


167

[7] V. V. Dmitriev, A. A. Senin, A. A. Soldatov, and A. N. Yudin. Polar phase of

superfluid 3 He in anisotropic aerogel. Phys. Rev. Lett., 115:165304, 2015.

[8] Asadchikov, V.E., Askhadullin, R.S., Volkov, V.V. et al. Structure and properties

of “nematically ordered” aerogels. JETP Letters, 101(8):556–561, Apr 2015.

[9] E. P. Wigner. The unreasonable e↵ectiveness of mathematics in the natural sciences.

Communications on Pure and Applied Mathematics, 13:1–14, 1960.

[10] A. J Leggett. Can a solid be ”superfluid”? Physical Review Letters, 25(22):1543–

1546, Nov.

[11] A. J. Leggett. Interpretation of Recent Results on 3 He below 3 mK: A New Liquid

Phase? Phys. Rev. Lett., 18, 1972.

[12] E. Schrödinger. An undulatory theory of the mechanics of atoms and molecules.

Phys. Rev., 28:1049–1070, Dec 1926.

[13] London, F. On the Bose-Einstein Condensation. Phys. Rev., 54:947–954, Dec 1938.

[14] L. Onsager. Statistical Hydrodynamics. Il Nouvo Cimento, 6:279–287, 1949.

[15] R. P. Feynman. The application of quantum mechanics to liquid helium. Vol. 1,

1955.

[16] Yang, C. N. Concept of O↵-Diagonal Long-Range Order and the Quantum Phases

of Liquid He and of Superconductors. Rev. Mod. Phys., 34:694–704, Oct 1962.


168

[17] Feynman, R. P. Atomic Theory of Liquid Helium Near Absolute Zero. Phys. Rev.,

91:1301–1308, Sep 1953.

[18] Feynman, R. P. Atomic Theory of the Transition in Helium. Phys. Rev., 91:1291–

1301, Sep 1953.

[19] Feynman, R. P. Atomic Theory of the Two-Fluid Model of Liquid Helium. Phys.

Rev., 94:262–277, Apr 1954.

[20] R. P. Feynman, F. L. Vernon, and R. W. Hellwarth. Geometrical Representation of

the Schrödinger Equation for Solving Maser Problems. J. Appl. Phys., 28:49, 1957.

[21] Kapitza, P. Viscosity of Liquid Helium below the - Point. Nature, 141:74, 1938.

[22] ALLEN, J., MISENER, A. Flow Phenomena in Liquid Helium II. Nature, 142:643–

644, 1938.

[23] D. D. Oshero↵, W.J. Gully, R. C. Richardson, and D. M. Lee. New Magnetic Phe-

nomena in Liquid 3 He below 3 mK. Phys. Rev. Lett., 29:920–923, 1972.

[24] O.T. Ikkala, G.E. Volovik, P.J. Hakonen, Yu Bunkov, S.T. Islander, and G.A. Khar-

odze. NMR in Rotating Superfluid 3 He-B, volume 35. American Institute of Physics,

1982.

[25] Hakonen, P. J. and Ikkala, O. T. and Islander, S. T. and Lounasmaa, O. V. and

Volovik, G. E. NMR Experiments on Rotating Superfluid 3 He-A and 3 He-B and


169

their Theoretical Interpretation. Journal of Low Temperature Physics, 53(3):425–

476, Nov 1983.

[26] J. P. Pekola, J. T. Simola, K. K. Nummila, O. V. Lounasmaa, and R. E. Packard.

Persistent-Current Experiments on Superfluid 3 He-B and 3 He-A. Phys. Rev. Lett.,

53(1):70–73, Jul 1984.

[27] Richard E. Packard and T. M. Sanders. Observations on single vortex lines in ro-

tating superfluid helium. Phys. Rev. A, 6:799–807, Aug 1972.

[28] Autti, S. and Dmitriev, V. V. and Mäkinen, J. T. and Soldatov, A. A. and Volovik,

G. E. and Yudin, A. N. and Zavjalov, V. V. and Eltsov, V. B. Observation of

Half-Quantum Vortices in Topological Superfluid 3 He. Phys. Rev. Lett., 117:255301,

2016.

[29] T W B Kibble. Topology of cosmic domains and strings. Journal of Physics A:

Mathematical and General, 9(8):1387–1398, aug 1976.

[30] Zurek, W. Cosmological experiments in superfluid helium? Nature, pages 505–508,

1985.

[31] W. H. Zurek. Cosmological Experiments in Condensed Matter Systems. Phys. Rep.,

276:177, 1996.

[32] Bardeen, J. and Cooper, L. N. and Schrie↵er, J. R. Theory of Superconductivity.

Phys. Rev., 108:1175–1204, Dec 1957.


170

[33] Cooper, Leon N. Bound Electron Pairs in a Degenerate Fermi Gas. Phys. Rev.,

104:1189–1190, Nov 1956.

[34] G. Eilenberger. Transformation of Gor’kov’s Equation for Type II Superconductors

into Transport-Like Equations. Z. Physics, 1968.

[35] Larkin, A. I. and Ovchinnikov, Yu. N. Quasiclassical Method in the Theory of

Superconductivity. Soviet Journal of Experimental and Theoretical Physics, 28:1200,

June 1969.

[36] Gor’kov, L P. Microscopic derivation of the Ginzburg–Landau equations in the

theory of superconductivity. Sov. Phys. - JETP (Engl. Transl.); (United States),

9:6, 1 1959.

[37] Yoichiro Nambu. Quasi-Particles and Gauge Invariance in the Theory of Supercon-

ductivity. Phys. Rev., 117:648–663, Feb 1960.

[38] J. A. Sauls. Lectures at Northwestern University. 2017.

[39] J. W. Serene and D. Rainer. The Quasiclassical Approach to 3 He. Phys. Rep.,

101:221, 1983.

[40] L. V. Keldysh. Diagram technique for nonequilibrium processes. Sov. Phys. JETP,

20:1018, 1965.

[41] Baym, Gordon and Kadano↵, Leo P. Conservation Laws and Correlation Functions.

Phys. Rev., 124:287–299, Oct 1961.


171

[42] Baym, Gordon. Self-Consistent Approximations in Many-Body Systems. Phys. Rev.,

127:1391–1401, Aug 1962.

[43] Y. Nambu. Fermion-Boson relations in BCS-type theories. Physica D: Nonlinear

Phenomena, 15(1–2):147 – 151, 1985.

[44] D. Rainer, J. A. Sauls, and D. Waxman. Current carried by bound states of a

superconducting vortex. Phys. Rev. B, 54:10094–10106, Oct 1996.

[45] Hao Wu and J. A. Sauls. Majorana excitations, spin and mass currents on the

surface of topological superfluid 3 He-B. Phys. Rev. B, 88:184506, 2013.

[46] T. Matsubara. A new approach to quantum-statistical mechanics. Prog. Theor.

Phys., 14:351–378, 1955.

[47] Riccati, Jacopo. Animadversiones in aequationes di↵erentiales secundi gradus. Ac-

torum Eruditorum, quae Lipsiae publicantur, 8:66–73, 1724.

[48] Ueki, Hikaru and Morita, Hiroki and Ohuchi, Marie and Kita, Takafumi. Drastic

enhancement of the thermal Hall angle in a d-wave superconductor. Phys. Rev. B,

101:184518, May 2020.

[49] Shelankov, A. and Ozana, M. Quasiclassical theory of superconductivity: A

multiple-interface geometry. Phys. Rev. B, 61:7077–7100, Mar 2000.

[50] J. A. Sauls. Broken Symmetry and Non-Equilibrium Superfluid 3 He, pages 239–265.

Elsievier Science Publishers, Amsterdam, 2000.


172

[51] A. J. Leggett. Microscopic Theory of NMR in an Anisotropic Superfluid 3 He-A.

Phys. Rev. Lett., 31:352, 1973.

[52] A. J. Leggett. NMR lineshifts and spontaneously broken spin-orbit symmetry - 1.

general concepts. J. Phys C, 6:3187, 1973.

[53] Balian, R. and Werthamer, N. R. Superconductivity with Pairs in a Relative p

Wave. Phys. Rev., 131:1553–1564, Aug 1963.

[54] P. W. Anderson and P. Morel. Generalized Bardeen-Cooper-Schrie↵er States and

the Proposed Low-Temperature Phase of 3 He. Phys. Rev., 123:1911, 1961.

[55] P. W. Anderson and W. Brinkman. Anisotropic Superfluidity in 3 He-A: A Possible

Interpretation of Its Stability as a Spin-Fluctuation E↵ect. Phys. Rev. Lett., 30:1108,

1973.

[56] V. J. Emery and A. M. Sessler. Possible Phase Transition in Liquid 3 He. Phys. Rev.,

119:43, 1960.

[57] V. J. Emery. Fluctuations above the superfluid transition in liquid 3 He. J. Low

Temp. Phys., 22:467, 1976.

[58] Ginzburg, V.L., Landau, L.D. On the theory of superconductivity. Zh. Eksp. Teor.

Fiz, 20:1064, 1950.

[59] N. D. Mermin and G. Stare. Ginzburg-landau approach to l 6= 0 pairing. Phys. Rev.

Lett., 30:1135, 1973.


173

[60] Ambegaokar, V. and deGennes, P. G. and Rainer, D. Landau-Ginsburg equations

for an anisotropic superfluid. Phys. Rev. A, 9:2676–2685, Jun 1974.

[61] Ho, Tin-Lun. Coreless vortices in superfluid 3 He-A: Topological structure, nucle-

ation, and the screening e↵ect. Phys. Rev. B, 18:1144–1153, Aug 1978.

[62] Maki, Kazumi and Kumar, Pradeep. Composite solitons and magnetic resonances

in superfluid 3 He-A. Phys. Rev. B, 16:182–190, Jul 1977.

[63] K. Maki. Solitons in Condensed Matter Physics. Springer Verlag, Berlin, 1978).

[64] E.V. Thuneberg. Ginzburg-Landau theory of vortices in superfluid 3 He-B. Phys.

Rev. B, 36:3583–3597, 1987.

[65] Regan, Robert C. and Wiman, J. J. and Sauls, J. A. Vortex phase diagram of

rotating superfluid 3 He B. Phys. Rev. B, 101:024517, Jan 2020.

[66] P. Sharma and J. A. Sauls. Investigation of quantized circulation in superfluid 3He-

B. Journal of Low Temperature Physics, 91:315–339, 1993.

[67] D. Vollhardt and P. Wölfle. The Superfluid Phases of Helium 3. Taylor & Francis,

New York, 1990.

[68] M. M. Salomaa and G. E. Volovik. Quantized vortices in superfluid he3. Rev. Mod.

Phys., 59(3):533–613, 1987.

[69] O. V. Lounasmaa and E. Thuneberg. Vortices in rotating superfluid 3 He. Proc. Nat.

Acad. Sci., 96, 1999.


174

[70] Salomaa, M. M. and Volovik, G. E. Vortices with Ferromagnetic Superfluid Core in


3
He-B. Phys. Rev. Lett., 51:2040–2043, Nov 1983.

[71] Thuneberg, E. V. Nucleation of superconductivity around weakly scattering defects.

Journal of Low Temperature Physics, 62(1):27–37, Jan 1986.

[72] Hakonen, P. J. and Krusius, M. and Salomaa, M. M. and Simola, J. T. and Bunkov,

Yu. M. and Mineev, V. P. and Volovik, G. E. Magnetic Vortices in Rotating Super-

fluid 3 He-B. Phys. Rev. Lett., 51:1362–1365, Oct 1983.

[73] J J Wiman and J A Sauls. Superfluid phases of 3 He in nanoscale channels. Phys.

Rev. B, 92(14):144515, 2015.

[74] Ohmi, Tetsuo and Tsuneto, Toshihiko and Fujita, Toshimitsu. Core Structure of

Vortex Line in 3He-B. Progress of Theoretical Physics, 70(3):647–653, 1983.

[75] M. Salomaa and G. E. Volovik. Half-quantum vortices in superfluid 3 He-a. Phys.

Rev. Lett., 55:1184–1187, Sep 1985.

[76] E. V. Thuneberg. Hydrostatic theory of superfluid 3 He-B. J. Low. Temp. Phys.,

122:657, 2019.

[77] A. Layzer and D. Fay. Spin-Fluctuation Exchange Mechanism for P-wave Pairing

in Liquid 3 He. Int. J. Magn., 1:135, 1971.

[78] D. Rainer and J. W. Serene. Free Energy of Superfluid 3 He. Phys. Rev. B, 13:4745,

1976.
175

[79] J. A. Sauls and J. W. Serene. Potential Scattering Models for the Quasiparticle

Interactions in Liquid 3 He. Phys. Rev. B, 24:183, 1981.

[80] J. A. Sauls and J. W. Serene. Higher-Order Strong Coupling E↵ects in Superfluid


3
He. Physica, B108:1137, 1981.

[81] Jorge Nocedal. Updating Quasi-Newton Matrices with Limited Storage. Mathemat-

ics of Computation, 35(151):773–782, 1980.

[82] J. Nocedal and S. J. Wright. Numerical Optimization. 2nd Ed., 2006.

[83] J. K. Viljas and E. V. Thuneberg. Equilibrium simulations of 2D weak links in

p-wave superfluids. J. Low. Temp. Phys., 129:423, 2002.

[84] Hasegawa, Yasumasa. On Vortex in Superfluid 3 He-B. Progress of Theoretical

Physics, 73(5):1258–1260, 1985.

[85] Silaev, M. A. and Thuneberg, E. V. and Fogelström, M. Lifshitz Transition in the

Double-Core Vortex in 3 He B. Phys. Rev. Lett., 115:235301, Dec 2015.

[86] Barnett, S. J. Magnetization by Rotation. Phys. Rev., 6:239–270, Oct 1915.

[87] Wiman, J. J. and Sauls, J. A. Spontaneous Helical Order of a Chiral p-Wave Su-

perfluid Confined in Nanoscale Channels. Phys. Rev. Lett., 121:045301, Jul 2018.

[88] Oleksii Shevtsov and J. A. Sauls. Electron Bubbles and Weyl Fermions in Chiral

Superfluid 3 He-A. Phys. Rev. B, 94:064511, 2016.


176

[89] Gary A. Williams and Richard E. Packard. Photographs of Quantized Vortex Lines

in Rotating He II. Phys. Rev. Lett., 33:280–283, Jul 1974.

[90] L. D. Landau and I. M. Lifshitz. Statistical Physics, volume 2nd ed., Vol. 5, Course

in Theoretical Physics. Pergamon Press, New York, 1969.

[91] J. A. Sauls. Anisotropic Pairing in Neutron Stars and Liquid 3 He, Ph.D. Thesis.

State University of New York at Stony Brook, (1980).

[92] T. Perry, K. DeConde, J.A. Sauls, and D.L. Stein. Evidence for Magnetic Coupling

in the Thermal Boundary Resistance Between Liquid 3 He and Platinum. Phys. Rev.

Lett., 48:1831, 1982.

[93] H Choi, J. P. Davis, J. Pollanen, T.M. Haard, and W.P. Halperin. Strong cou-

pling corrections to the Ginzburg-Landau theory of superfluid 3 He. Phys. Rev. B,

75(17):174503, 2007.

[94] Sauls, J. A. and Mizushima, Takeshi. On the Nambu fermion-boson relations for

superfluid 3 He. Phys. Rev. B, 95:094515, Mar 2017.

[95] M. D. Nguyen, A. M. Zimmerman, and W. P. Halperin. Corrections to higgs mode

masses in superfluid 3 He from acoustic spectroscopy. Phys. Rev. B, 99:054510, Feb

2019.

[96] D. Vollhardt and P. Wölfle. The Superfluid Phases of Helium 3. Dover, New York,

2013.
177

[97] William P. Halperin, Jeevak M. Parpia, James A. Sauls . Superfluid helium-3 in

confined quarters. Physics Today, 71, 2018.

[98] W.P. Halperin. Superfluid 3 He in Aerogel. Annual Review of Condensed Matter

Physics, 10(1):155–170, 2019.

[99] J. A. Sauls. Viewpoint: Half-Quantum Vortices in Superfluid Helium. Physics 9,

148, 2016.

[100] Dmitriev, V.V., Krasnikhin, D.A., Mulders, N. et al. Orbital glass and spin glass

states of 3He-A in aerogel. JETP Letters, 91:599–606, 2010.

[101] Makinen, J. and Dmitriev, V. V. and Nissinen, J. T. and Rysti, J. and Volovik, G.

E. and Yudin, A. N. and Zhang, K. and V. B. Elstov. Half-quantum vortices and

walls bounded by strings in the polar-distorted phases of topological superfluid 3 He.

Nature Communications 10, 237, 2019.

[102] V.B. Eltsov and M. Krusius and G.E. Volovik. Vortex formation and dynamics in

superfluid 3He and analogies in quantum field theory. volume 15 of Progress in Low

Temperature Physics, pages 1 – 137. Elsevier, 2005.

[103] A del Campo and A Retzker and M B Plenio. The inhomogeneous Kibble–Zurek

mechanism: vortex nucleation during Bose–Einstein condensation. New Journal of

Physics, 13(8):083022, aug 2011.


178

[104] Chesler, Paul M. and Garcı́a-Garcı́a, Antonio M. and Liu, Hong. Defect Formation

beyond Kibble-Zurek Mechanism and Holography. Phys. Rev. X, 5:021015, May

2015.

[105] J. Rysti, S. Autti, G.E. Volovik, V.B. Eltsov. Kibble-Zurek creation of half-quantum

vortices under symmetry violating bias. arXiv:1906.11453, 2019.

[106] Dmitri A Ivanov. Non-Abelian statistics of half-quantum vortices in p-wave super-

conductors. Physical Review Letters, 86(2):268, 2001.

[107] J. E. Baumgardner and D. D. Oshero↵. Phase Diagram of Superfluid 3 He in 99.3%

Porosity Aerogel. Phys. Rev. Lett., 93:155301, 2004.

[108] R Sh Askhadullin, V V Dmitriev, D A Krasnikhin, P N Martynov, A A Osipov, A A

Senin, and A N Yudin. Phase diagram of superfluid 3 He in “nematically ordered”

aerogel. JETP Lett., 95(6):326–331, 2012.

[109] H. C. Choi, A. J. Gray, C. L. Vicente, J. S. Xia, G. Gervais, W. P. Halperin,

N. Mulders, and Y. Lee. A1 and a2 transitions in superfluid 3 He in 98% porosity

aerogel. Phys. Rev. Lett., 93(14):145302, 2004.

[110] H. Choi, K. Yawata, T.M. Haard, J.P. Davis, G. Gervais, N. Mulders, P. Sharma,

J.A. Sauls, and W.P. Halperin. Specific heat of disordered superfluid 3 He. Phys.

Rev. Lett., 93(14):145301, 2004.


179

[111] J Pollanen, J. I. A. Li, C. A. Collett, W. J. Gannon, and W P Halperin. Identification

of Superfluid Phases of He-3 in Uniformly Isotropic 98.2% Aerogel. Phys. Rev. Lett.,

107(19):195301, 2011.

[112] J Pollanen, J. I. A. Li, C. A. Collett, W. J. Gannon, W P Halperin, and J A Sauls.

New chiral phases of superfluid He-3 stabilized by anisotropic silica aerogel. Nat.

Phys., 8(4):317–320, 2012.

[113] J. A. Sauls and P. Sharma. Impurity E↵ects on the A1 -A2 Splitting of Superfluid
3
He in Aerogel. Phys. Rev. B, 68:224502, 2003.

[114] J. A. Sauls. Chiral phases of superfluid 3 He in an anisotropic medium. Phys. Rev.

B, 88:214503, 2013.

[115] Kazushi Aoyama and Ryusuke Ikeda. Pairing states of superfluid 3 He in uniaxially

anisotropic aerogel. Phys. Rev. B, 73(6), 2006.

[116] Sauls, James A and Bunkov, Yu M. and E. Collin and H. Godfrin and P. Sharma.

Magnetization and spin di↵usion of liquid He3 in aerogel. Physical Review B, 72(2),

7 2005.

[117] J. Jang, D. G. Ferguson, V. Vakaryuk, R. Budakian, S. B. Chung, P. M. Goldbart,

and Y. Maeno. Observation of Half-Height Magnetization Steps in Sr2 RuO4 . Science,

331(6014):186–188, 2011.

[118] G. E. Volovik and V. P. Mineev. Line and Point Singularities of Superfluid 3 He.

Sov. Phys. JETP Lett., 24(11):561–563, 1976. [Pis’ma ZhETF, 24, 605-608 (1976)].
180

[119] Hisamitsu, Tomohiro and Tange, Masaki and Ikeda, Ryusuke. Impact of strong

anisotropy on the phase diagram of superfluid 3 He in aerogels. Phys. Rev. B,

101:100502, Mar 2020.

[120] I. A. Fomin. Robust Superfluid Phases 3 He in Aerogel. J. Low Temp. Phys., 134:769,

2004.

[121] I. A. Fomin. E↵ect of random anisotropy on the shift of the nmr frequency in the

polar phase of superfluid 3 he. JETP Letters, 109(5):325–330, 2019.

[122] Hu, Chia-Ren and Maki, Kazumi. Satellite magnetic resonances of a bound pair

of half-quantum vortices in rotating superfluid 3 He-A. Phys. Rev. B, 36:6871–6880,

Nov 1987.

[123] Mineev, V. P. Half-Quantum Vortices in Polar Phase of Superfluid 3 He. J. Low

Temp. Phys., 177(1):48–58, 2014.

[124] V. P. Mineev. NMR Properties of the Polar Phase of Superfluid 3 He. J. Low Temp.

Phys., 184:1007–1014, 2016.

[125] Nagamura, Natsuo and Ikeda, Ryusuke. Stability of half-quantum vortices in equal-

spin pairing states of 3 He. Phys. Rev. B, 98:094524, Sep 2018.

[126] Tange, Masaki and Ikeda, Ryusuke. Half-quantum vortex pair in the polar-distorted

B phase of superfluid 3 He in aerogels. Phys. Rev. B, 101:094512, Mar 2020.


181

[127] Pye Ton How, Sung-Kit Yip. Half-quantum vortices in nematic superconductors.

arXiv:2005.03366, 2020.

[128] Chung, Suk Bum and Bluhm, Hendrik and Kim, Eun-Ah. Stability of Half-Quantum

Vortices in px + ipy Superconductors. Phys. Rev. Lett., 99:197002, Nov 2007.

[129] Nagai, Yuki and Kato, Yusuke. Quasiparticle Bound States around Fractional

Vortices in s-wave Superconductor. Journal of the Physical Society of Japan,

88(5):054707, 2019.

[130] Ichioka, Masanori and Hasegawa, Akiko and Machida, Kazushige. Field depen-

dence of the vortex structure in d-wave and s-wave superconductors. Phys. Rev.

B, 59:8902–8916, Apr 1999.

[131] Ichioka, Masanori and Machida, Kazushige. Field dependence of the vortex structure

in chiral p-wave superconductors. Phys. Rev. B, 65:224517, Jun 2002.

[132] Masaki, Yusuke and Mizushima, Takeshi and Nitta, Muneto. Microscopic descrip-

tion of axisymmetric vortices in 3 P2 superfluids. Phys. Rev. Research, 2:013193, Feb

2020.

[133] White, J. S. and Hinkov, V. and Heslop, R. W. and Lycett, R. J. and Forgan,

E. M. and Bowell, C. and Strässle, S. and Abrahamsen, A. B. and Laver, M. and

Dewhurst, C. D. and Kohlbrecher, J. and Gavilano, J. L. and Mesot, J. and Keimer,

B. and Erb, A. Fermi Surface and Order Parameter Driven Vortex Lattice Structure

Transitions in Twin-Free YBa2 Cu3 O7 . Phys. Rev. Lett., 102:097001, Mar 2009.
182

[134] Avers, K.E., Gannon, W.J., Kuhn, S.J. et al. Broken time-reversal symmetry in the

topological superconductor UPt3. Nature Physics, 2020.

[135] Askhadullin, R. Sh. and Dmitriev, V. V. and Martynov, P. N. and Osipov, A. A. and

Senin, A. A. and Yudin, A. N. Anisotropic 2D Larkin-Imry-Ma state in the polar

distorted ABM phase of 3He in a “nematically ordered” aerogel. JETP Letters,

100(10):662–668, Jan 2015.

[136] Eschrig, M. and Sauls, J. A. and Rainer, D. Electromagnetic response of a vortex

in layered superconductors. Phys. Rev. B, 60:10447–10454, Oct 1999.

[137] Kato, Yusuke. Phase-Sensitive Impurity E↵ects in Vortex Core of Moderately Clean

Chiral Superconductors. Journal of the Physical Society of Japan, 69(10):3378–3386,

2000.

[138] Kato, Yusuke and Hayashi, Nobuhiko. Numerical Study of Impurity E↵ects on

Quasiparticles within S-wave and Chiral P-wave Vortices. Journal of the Physical

Society of Japan, 71(7):1721–1727, 2002.

[139] M Eschrig and J A Sauls. Charge dynamics of vortex cores in layered chiral triplet

superconductors. New Journal of Physics, 11(7):075009, jul 2009.

[140] J. A. Sauls and M. Eschrig. Vortices in chiral, spin-triplet superconductors and

superfluids. New Journal of Physics, 11(7):075008, jul 2009.

[141] Masaki, Yusuke. Vortex charges and impurity e↵ects based on quasiclassical theory

in an s-wave and a chiral p-wave superconductor. Phys. Rev. B, 99:054512, 2019.


183

[142] J. Nocedal and S. J. Wright. Numerical Optimization, 2nd ed. Springer, Berlin,

2006.

[143] M.M. Salomaa and G.E. Volovik. Symmetry and Structure of Quantized Vortices

in Superfluid 3 He-B. Phys. Rev. Lett., 31:203–227, 1985.

[144] J. A. Sauls and J. W. Serene. Interaction E↵ects on the Zeeman Splitting of Col-

lective Modes in Superfluid 3 He-B. Phys. Rev. Lett., 49:1183, 1982.

[145] P. Sharma and J. A. Sauls. Magnetic Susceptibility of the Balian-Werthamer Phase

of 3 He in Aerogel. J. Low Temp. Phys., 125:115, 2001.

[146] A.J Leggett. The Spin Dynamics of an Anisotropic Fermi Supefluid 3 He. Ann. Phys.,

85(1):11–55, 1974.

[147] V V. Zavjalov. Linear NMR in the Polar Phase of 3He in Aerogel. JETP Letters,

09 2018.

[148] Samuli Autti. Ph.D. Thesis. Aalto University, 2017.

[149] Kopnin, N. B. and Kravtsov, V. E. Forces acting on vortices moving in a pure type

II superconductor. Soviet Journal of Experimental and Theoretical Physics, 44:861,

October 1976.

[150] Yusuke Kato and Chun-Kit Chung. Nature of driving force on an isolated mov-

ing vortex in dirty superconductors. Journal of the Physical Society of Japan,

85(3):033703, 2016.
184

[151] Embon, L., Anahory, Y., Jelić, Ž. et al. Imaging of super-fast dynamics and flow

instabilities of superconducting vortices. Nature Comm., 8:85, 2017.

[152] Dobrovolskiy, O.V., Vodolazov, D.Y., Porrati, F. et al. Ultra-fast vortex motion in

a direct-write Nb-C superconductor. Nature Comm., 1:3291, 2020.

[153] Kogan, V. G. and Prozorov, R. Interaction between moving Abrikosov vortices in

type-II superconductors. Phys. Rev. B, 102:024506, Jul 2020.

[154] Eskildsen, M. R. and Abrahamsen, A. B. and López, D. and Gammel, P. L. and

Bishop, D. J. and Andersen, N. H. and Mortensen, K. and Canfield, P. C. Flux Line

Lattice Reorientation in the Borocarbide Superconductors with H k a. Phys. Rev.

Lett., 86:320–323, Jan 2001.

[155] Okuma, S. and Kashiro, K. and Suzuki, Y. and Kokubo, N. Order-disorder transi-

tion of vortex matter in a-Mox Ge1 x films probed by noise. Phys. Rev. B, 77:212505,

Jun 2008.

[156] Okuma, S. and Imaizumi, H. and Shimamoto, D. and Kokubo, N. Quantum melting

and lattice orientation of driven vortex matter. Phys. Rev. B, 83:064520, Feb 2011.

[157] Das, P. and Rastovski, C. and O’Brien, T. R. and Schlesinger, K. J. and Dewhurst,

C. D. and DeBeer-Schmitt, L. and Zhigadlo, N. D. and Karpinski, J. and Eskildsen,

M. R. Observation of Well-Ordered Metastable Vortex Lattice Phases in Supercon-

ducting MgB2 Using Small-Angle Neutron Scattering. Phys. Rev. Lett., 108:167001,

Apr 2012.
185

[158] Hirano, Tomoya and Takamoril, Kenta and Ichioka, Masanori and Machida,

Kazushige. Rotation of Triangular Vortex Lattice in the Two-Band Superconductor

MgB2. Journal of the Physical Society of Japan, 82(6):063708, 2013.

[159] Masaki, Yusuke and Kato, Yusuke. Impurity E↵ects on Bound States in Vortex Core

of Topological s-Wave Superconductor. Journal of the Physical Society of Japan,

84(9):094701, 2015.

[160] Masaki, Yusuke and Kato, Yusuke. Impurity E↵ects on Caroli–de Gennes–Matricon

Mode in Vortex Core in Superconductors. Journal of the Physical Society of Japan,

85(1):014705, 2016.

[161] Yusuke Kato. Charging e↵ect on the hall conductivity of single vortex in type ii

superconductors. Journal of the Physical Society of Japan, 68(12):3798–3801, 1999.

[162] Kumagai, Ken-ichi and Nozaki, Koji and Matsuda, Yuji. Charged vortices in high-

temperature superconductors probed by NMR. Phys. Rev. B, 63:144502, Mar 2001.

[163] Mounce, A., Oh, S., Mukhopadhyay, S. et al. Charge-induced vortex lattice insta-

bility. Nature Physics, pages 125–128, 2011.

[164] Ueki, Hikaru and Kohno, Wataru and Kita, Takafumi. Vortex-Core Charging Due

to the Lorentz Force in a d-Wave Superconductor. Journal of the Physical Society

of Japan, 85(6):064702, 2016.


186

[165] Ueki, Hikaru and Ohuchi, Marie and Kita, Takafumi. Charging in a Superconducting

Vortex Due to the Three Force Terms in Augmented Eilenberger Equations. Journal

of the Physical Society of Japan, 87(4):044704, 2018.

[166] Kohno, Wataru and Ueki, Hikaru and Kita, Takafumi. Hall E↵ect in the Vortex

Lattice of d-Wave Superconductors with Anisotropic Fermi Surfaces. Journal of the

Physical Society of Japan, 86(2):023702, 2017.

[167] Ohuchi, Marie and Ueki, Hikaru and Kita, Takafumi. Charging due to Pair-Potential

Gradient in Vortex of Type-II Superconductors. Journal of the Physical Society of

Japan, 86(7):073702, 2017.

[168] Marie Ohuchi, Hikaru Ueki, Takafumi Kita. Anomalous Hall e↵ect in the Abrikosov

lattice of type-II superconductors. arXiv:2011.04856, 2020.

[169] Y. Tanaka. Soliton in Two-Band Superconductor. Phys. Rev. Lett., 88:017002, Dec

2001.

[170] Babaev, Egor. Vortices with Fractional Flux in Two-Gap Superconductors and in

Extended Faddeev Model. Phys. Rev. Lett., 89:067001, Jul 2002.

[171] Runge, C. Ueber die numerische Auflösung von Di↵erentialgleichungen. Math. Ann.,

46:167–178, 1895.

[172] Kutta, Martin. Beitrag zur näherungsweisen Integration totaler Di↵erentialgleichun-

gen. Zeitschrift für Mathematik und Physik, 46:435–453, 1901.


187

[173] Nagai, Yuki and Kato, Yusuke and Hayashi, Nobuhiko. Analytical Result on Elec-

tronic States around a Vortex Core in a Noncentrosymmetric Superconductor. Jour-

nal of the Physical Society of Japan, 75(4):043706, 2006.

[174] Nagai, Yuki and Ueno, Yosuke and Kato, Yusuke and Hayashi, Nobuhiko. Analytical

Formulation of the Local Density of States around a Vortex Core in Unconventional

Superconductors. Journal of the Physical Society of Japan, 75(10):104701, 2006.

[175] Mel’nikov, A. S. and Ryzhov, D. A. and Silaev, M. A. Local density of states around

single vortices and vortex pairs: E↵ect of boundaries and hybridization of vortex

core states. Phys. Rev. B, 79:134521, Apr 2009.

[176] Nagai, Yuki. N-independent Localized Krylov–Bogoliubov-de Gennes Method:

Ultra-fast Numerical Approach to Large-scale Inhomogeneous Superconductors.

Journal of the Physical Society of Japan, 89(7):074703, 2020.

You might also like