p s remnants: rotational and magnetic e their progenitors ......yevgeni kissin doctor of philosophy...

219
Properties of Stellar Remnants:Rotational and Magnetic Evolution of T heir Progenitors by Yevgeni Kissin A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Astronomy & Astrophysics University of Toronto Copyright c 2017 by Yevgeni Kissin

Upload: others

Post on 04-Mar-2021

6 views

Category:

Documents


0 download

TRANSCRIPT

  • Properties of Stellar Remnants: Rotational and MagneticEvolution of Their Progenitors

    by

    Yevgeni Kissin

    A thesis submitted in conformity with the requirementsfor the degree of Doctor of Philosophy

    Graduate Department of Astronomy & AstrophysicsUniversity of Toronto

    Copyright c© 2017 by Yevgeni Kissin

  • Abstract

    Properties of Stellar Remnants: Rotational and Magnetic Evolution of Their

    Progenitors

    Yevgeni Kissin

    Doctor of Philosophy

    Graduate Department of Astronomy & Astrophysics

    University of Toronto

    2017

    The rotation and magnetic evolution of stars is examined in detail. Advection of

    angular momentum by convective plumes is a source of differential rotation, which

    both reduces the angular momentum loss to stellar winds and spins up the inner

    regions. This spin up not only affects the rotation of the stellar remnant, but also

    facilitates the winding of magnetic field lines necessary to amplify a seed field. In

    radiative regions we demonstrate that an embedded magnetic field can effectively mix

    angular momentum, as well as sustain coupling with adjacent convective zones.

    In addition to the resulting angular momentum transport model, we consider a

    mechanism for rotationally-induced winding of a seed magnetic field. When this

    mechanism operates at a convective-radiative boundary which moves in a Lagrangian

    sense towards the convection zone, net helicity is left behind in the newly radiative

    material.

    We employ this model of angular momentum transport and helicity deposition

    in stellar models of a wide range of masses. Our aim is to gauge the effect on the

    evolution of rotation and the impact on the properties of stellar remnants. In low-

    and intermediate-mass stars we find a quantitative agreement between the predicted

    and observed core and surface rotation rates, during both the sub-giant and helium-

    burning phases. Continuing the evolution to the ejection of the envelope, we are able

    ii

  • to reproduce the spin periods and magnetic fields of highly magnetic white dwarf

    remnants.

    We find similar success in massive stellar models, by being able to reproduce both

    the spin periods and magnetic fields of pulsars. And in progenitors of black holes

    with sufficiently high angular momentum pre-core-collapse, a portion of the infalling

    material can reach rotational support. This quenches the accretion onto the newly

    formed black hole, and may result in observable phenomena, such as high energy jets.

    iii

  • Dedication

    I would like to dedicate this dissertation to my soon-to-be-born first child. May it

    serve as an inspiration to aim high and never give up!

    iv

  • Acknowledgements

    I would like to thank several people who played a key role in my achievement:

    Christopher Thompson for his dedicated supervision over the past six years. I learned

    to pay close attention to details, to be thorough, and that always "physics works".

    My parents, Olga and Yakov Kissin, for all the sacrifices they made so their children can

    have a better and easier life than they did (as exemplified in their dual immigrations).

    And my wife, Anna Kissin, for supporting me both physically and emotionally

    throughout the entire duration of my PhD program. Me being in graduate school

    wasn’t easy for either of us, and it wasn’t her choice, but I couldn’t have done it

    without her.

    v

  • Contents

    1 Introduction 1

    1.1 Rotation and Angular Momentum in Stars . . . . . . . . . . . . . . . . . 3

    1.2 Angular Momentum Transport . . . . . . . . . . . . . . . . . . . . . . . . 5

    1.2.1 Low- and Intermediate-Mass Stars . . . . . . . . . . . . . . . . . . 7

    1.2.2 Angular Momentum Pumping by Convective Plumes and Rota-

    tional Coupling via Maxwell Stress . . . . . . . . . . . . . . . . . 9

    1.2.3 Massive Stars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

    1.2.4 Angular Momentum Transport by Convection and Maxwell

    Stresses in High-Mass Stars . . . . . . . . . . . . . . . . . . . . . . 15

    1.3 Binary Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    1.3.1 Low- and Intermediate-Mass Primaries . . . . . . . . . . . . . . . 18

    1.3.2 Massive Primaries . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

    1.4 Rotation and Magnetism of Stellar Remnants . . . . . . . . . . . . . . . . 22

    1.4.1 White Dwarfs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

    1.4.2 Neutron Stars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

    1.4.3 Black Holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    1.5 Plan of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    2 Rotation of Giant Stars 38

    2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

    2.1.1 Plan of the Chapter . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

    vi

  • 2.2 Pumping of Angular Momentum in Deep Convective Envelopes: Back-

    reaction from the Coriolis Force . . . . . . . . . . . . . . . . . . . . . . . . 40

    2.2.1 Extended Upflows and Downflows with Quasi-Geostrophic Bal-

    ance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

    2.2.2 Rapid Mixing Between Upflows and Downflows . . . . . . . . . 43

    2.2.3 Model Ω̄(r) Profiles . . . . . . . . . . . . . . . . . . . . . . . . . . 44

    2.2.4 Latitudinal Differential Rotation . . . . . . . . . . . . . . . . . . . 45

    2.2.5 Matching of Rotation between Core and Convective Envelope . . 47

    2.3 Comparison of 1.5 M� Model Star with Kepler Data . . . . . . . . . . . . 49

    2.3.1 Transition from Co & 1 to Co . 1 throughout Bulk of Convective

    Envelope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

    2.3.2 Surface Rotation of Subgiants: Angular Momentum Pumping

    Versus Magnetized Winds . . . . . . . . . . . . . . . . . . . . . . . 58

    2.4 Evolving Rotation Profile with Mass Loss and Interaction with a Planet 60

    2.4.1 Loss of Angular Momentum on the RGB and AGB . . . . . . . . 61

    2.4.2 Rotation Rate of the Inner Envelope and Core . . . . . . . . . . . 61

    2.5 Orbital Evolution of a Planetary Companion to a 1M� Star . . . . . . . 68

    2.5.1 Jupiter-mass Planet with ai = 1 AU . . . . . . . . . . . . . . . . . 70

    2.5.2 Other Planetary Configurations . . . . . . . . . . . . . . . . . . . 70

    2.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

    2.7 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

    2.7.1 Normalization of Convective Quadrupole During Giant Evolution 77

    3 Spin and Magnetism of White Dwarfs 79

    3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

    3.1.1 Plan of the Chapter . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

    3.2 Rotation and Magnetism in a Rapidly Evolving Giant Star . . . . . . . . 82

    3.2.1 Combined Effect of Convection and Poloidal Magnetic Field on

    Angular Momentum Redistribution . . . . . . . . . . . . . . . . . 82

    vii

  • 3.2.2 Magnetic Helicity Growth in Giant Cores . . . . . . . . . . . . . . 84

    3.2.3 Relative Importance of Different Convective Episodes . . . . . . 88

    3.3 Magnetic Field Amplification Near Core-Envelope Boundary . . . . . . 93

    3.3.1 Critical Co for Dynamo Action: Effect of an Intense Radiation

    Flux in the Tachocline . . . . . . . . . . . . . . . . . . . . . . . . . 93

    3.3.2 Magnetic Buoyancy in the Tachocline . . . . . . . . . . . . . . . . 95

    3.3.3 Seeding a Toroidal Magnetic Field in the Tachocline . . . . . . . 97

    3.4 White Dwarf Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . 101

    3.4.1 Accumulation of Magnetic Helicity in a Radiative Core Relaxing

    to Solid Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

    3.4.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

    3.5 White Dwarf Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

    3.5.1 Dependence of White Dwarf Spin on Coriolis Parameter for

    Core-Envelope Decoupling . . . . . . . . . . . . . . . . . . . . . . 114

    3.5.2 White Dwarf Spin in Case of Continued Core-Envelope Coupling114

    3.5.3 Minimal Magnetic Flux Enforcing Strong Core-Envelope Coupling118

    3.6 Magnetic Field Emergence and Decay . . . . . . . . . . . . . . . . . . . . 122

    3.7 Summary and Comparison with Alternative Theoretical Approaches . . 128

    3.7.1 Some Outstanding Issues . . . . . . . . . . . . . . . . . . . . . . . 131

    3.7.2 Competing effects of a collapsing low-mass (∼ 10−4 M�) AGB

    envelope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

    3.7.3 Impulsive Dynamo Amplification in a Merger? . . . . . . . . . . 136

    3.8 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

    3.8.1 Twisting of a Radial Magnetic Field . . . . . . . . . . . . . . . . . 138

    4 Rotation and Magnetism of Massive Stellar Cores 141

    4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

    4.1.1 Plan of the Chapter . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

    viii

  • 4.2 Stellar Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

    4.2.1 Building the Star . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

    4.2.2 Prescription for Numerical Evolution . . . . . . . . . . . . . . . . 148

    4.2.3 Angular Momentum Transport and Magnetic Field Evolution . . 150

    4.3 Rotational Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

    4.3.1 Initialization of the Rotation . . . . . . . . . . . . . . . . . . . . . 151

    4.3.2 Inward Pumping of Angular Momentum by Convection . . . . . 151

    4.3.3 Magnetic Angular Momentum Transport Limited by Kink Insta-

    bility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

    4.3.4 Rotational Evolution between Model Snapshots . . . . . . . . . . 154

    4.3.5 Some Basic Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

    4.3.6 Comparison with Rotationally Induced Mixing . . . . . . . . . . 157

    4.4 Magnetic Helicity Accumulation . . . . . . . . . . . . . . . . . . . . . . . 158

    4.4.1 Successive Convective Structures . . . . . . . . . . . . . . . . . . . 162

    4.5 Combined Magnetic and Rotational Evolution to Collapse . . . . . . . . 165

    4.5.1 Dependence on Effective Temperature . . . . . . . . . . . . . . . . 165

    4.5.2 Dependence on Stellar Mass . . . . . . . . . . . . . . . . . . . . . 167

    4.5.3 Neutron Star Remnants: Rotation and Dipole Magnetic Flux . . 167

    4.5.4 Black Hole Remnants: Mass and Spin . . . . . . . . . . . . . . . . 171

    4.6 Angular Momentum Injection from a Binary Companion . . . . . . . . . 173

    4.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177

    4.7.1 Implications for Growth or Decay of the Magnetic Field Post-

    Collapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

    4.8 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

    4.8.1 MESA inlist . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

    4.8.2 Mass loss prescription . . . . . . . . . . . . . . . . . . . . . . . . . 184

    4.8.3 Angular Momentum Transport by Winding a Poloidal Magnetic

    Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

    ix

  • 5 Conclusions & Future Work 188

    5.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

    5.2 Predictions and Observational Constraints . . . . . . . . . . . . . . . . . 191

    5.3 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193

    Bibliography 195

    x

  • List of Figures

    2.1 Illustration of the two rotation models proposed in this work. . . . . . . 47

    2.2 Flow speed of stellar material with respect to the core-envelope bound-

    ary, compared with convective speed in the inner envelope. . . . . . . . 50

    2.3 Rotation period at the base of the convective envelope, in a MZAMS =

    1.5M� model star, as a function of time during the RGB and early core

    He burning phases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

    2.4 Rotation period at the base of the convective envelope of 1.5 M� and

    M� models, for different rotation profiles. . . . . . . . . . . . . . . . . . 53

    2.5 Rotation period at the base of the convective envelope of 5 M� model,

    for different rotation rates. . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

    2.6 Profile of Coriolis parameter with radius during two phases of the

    evolution of our 1.5 M� model with initial vrot = 50 km s−1. . . . . . . . 56

    2.7 Profile of convective envelope during RGB expansion and AGB expan-

    sion in rotation model I. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

    2.8 Evolution of surface rotation rate post-MS of a 1.5M� model and com-

    parison with observations compiled by Schrijver & Pols (1993). . . . . . 59

    2.9 Evolution of spin angular momentum of 1 M� model star during RGB,

    core He burning and final AGB phases. . . . . . . . . . . . . . . . . . . . 64

    2.10 Evolution of the Coriolis parameter at the base of the convective envelope. 65

    2.11 Effect on the 5M� model star of the engulfment of a substellar companion. 66

    xi

  • 2.12 Same as right panel of Figure 2.11, but now with steeper inner rotation

    profile. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

    2.13 Stellar radius of a 1M� model post-MS, and the orbital semimajor axis

    of a planetary companion. . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

    2.14 Evolution of semimajor axis and eccentricity of a planetary companion

    around a 1M� model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

    2.15 Probability of engulfment of planets of various masses around a 1M�

    model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

    3.1 Evolution of the base of the convective envelope of a 1M� during the

    tips of the RGB and AGB. . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

    3.2 Evolution of the base of the convective envelope of a 5M� during the

    tips of the RGB and AGB. . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

    3.3 Illustration of the buoyancy of a magnetic flux tube in a tachocline layer. 97

    3.4 Relative buoyant speed of a magnetic flux tube in the tachocline layer

    during the tips of the RGB and AGB of a 1M� model. . . . . . . . . . . 98

    3.5 Thickness of a magnetized shell below the core-envelope boundary

    which remains in contact with the active dynamo region. . . . . . . . . . 99

    3.6 Minimal seed poloidal field for magnetic buoyancy to overcome the

    doward drift of material. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

    3.7 Magnetic helicity accumulation as a function of time in the 1M�. . . . . 107

    3.8 Comparison of the magnetic helicity generated via the two channels we

    consider, in the 5M� model. . . . . . . . . . . . . . . . . . . . . . . . . . . 108

    3.9 Dipole magnetic field left behind in the WD remnants of 1M� and 5M�

    models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

    3.10 Dipole magnetic field left behind in the WD remnants of a 5M� model

    as a function of companion mass. . . . . . . . . . . . . . . . . . . . . . . . 110

    3.11 Evolution of spin angular momentum of the 1M� model during the tip

    of the AGB. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

    xii

  • 3.12 Evolution of spin angular momentum of the 5M� model during the tip

    of the AGB. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

    3.13 Rotation period of the WD remnants of the 1M� and 5M� models. . . . 115

    3.14 Comparison of the minimum angular momentum to maintain cou-

    pling during the AGB wind phase of a 1M� and the stellar angular

    momentum in cases with planetary interaction. . . . . . . . . . . . . . . 116

    3.15 Total and core spin angular momentum of an AGB star during the con-

    traction of its hydrogen envelope and the dimensionless core moment

    of inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

    3.16 Comparison between different estimates of magnetic flux in order to

    motivate decoupling of core at end of AGB of 1M� model . . . . . . . . 121

    3.17 Comparison between different estimates of magnetic flux in order to

    decoupling of core at end of AGB of 5M� model . . . . . . . . . . . . . 122

    3.18 Ohmic diffusion timescale in magnetized layer of WD remnant of 1M�

    model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

    3.19 Ohmic diffusion timescale in magnetized layer of WD remnant of 5M�

    model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

    3.20 Ohmic diffusion and decay of magnetic flux . . . . . . . . . . . . . . . . 127

    3.21 Radius of surface and base of convective envelope versus residual

    hydrogen mass at the end of evolution of 1M� model . . . . . . . . . . . 135

    4.1 Determination of type of stellar remnant of our models based on com-

    pactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

    4.2 Convective structure of the 13M� model during both entire evolution

    and final year . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

    4.3 Convective structure of the 25M� model during final year of evolution 147

    4.4 Convective structure of the 25M� model during the pre-MS accretion

    phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

    xiii

  • 4.5 Kippenhahn diagram of zones which transport angular momentum and

    those in which it is frozen, during both entire evolution and final year

    of the 13M� model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

    4.6 Evolution of the internal angular momentum during the entire evolution

    of the 25M� model, using period of collapsed NS as a proxy . . . . . . 156

    4.7 Evolution of the internal angular momentum during the entire evolution

    of the 25M� model, using period of collapsed NS as a proxy, assuming

    no magnetic field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

    4.8 Comparison of timescales for angular momentum transport by different

    mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

    4.9 Magnetic flux contribution by different phases of the evolution of the

    13M� model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

    4.10 Magnetic flux contribution by different phases of the evolution of the

    25M� model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

    4.11 Magnetic flux contribution by different phases of the evolution of the

    two 40M� models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

    4.12 Comparison of the angular momentum loss to a wind for the four stellar

    models we consider . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

    4.13 Specific angular momentum profiles in the 13 M� and 25 M� models

    at four separate times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168

    4.14 Evolution of the internal angular momentum during the final year of

    evolution of the 13M� and 25M� models, using period of collapsed NS

    as a proxy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169

    4.15 Evolution of the internal angular momentum during the final year of

    evolution of the two 40M� models, using period of collapsed NS as a

    proxy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170

    xiv

  • 4.16 Specific angular momentum of the two isolated 40 M� models right

    before core collapse, compared with the corresponding specific angular

    of a stable orbit around a collapse BH . . . . . . . . . . . . . . . . . . . . 172

    4.17 Stellar radius and mass as a function of age for both 40 M� models . . . 175

    4.18 Impact of external torque from a companion on the angular momentum

    stored in the evolving core of the 13M� and 25M� models . . . . . . . . 176

    4.19 Impact of external torque from a companion on the angular momentum

    stored in the evolving core of the low metallicity 40M� model . . . . . . 177

    4.20 Specific angular momentum of the low metallicity 40 M� model right

    before core collapse, after experiencing an external torque from a com-

    panion, compared with the corresponding specific angular of a stable

    orbit around a collapse BH . . . . . . . . . . . . . . . . . . . . . . . . . . 178

    4.21 Spin angular momentum of the collapsed BH in the low metallicity

    40M� model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179

    xv

  • Chapter 1

    Introduction

    Stars are the building blocks of the universe. They produce the majority of radiative

    energy and are responsible for the chemical enrichment of the universe. It is therefore

    vital to understand the properties of stars, to be able to explain their impact on their

    surroundings. Stars illuminate and heat up the gas around them, and those with

    effective temperatures Teff & 30, 000K can even impact the formation of other stars in

    the same giant molecular cloud in which they formed. Once stars exhaust their main

    source of fuel, hydrogen, they expand to giant dimensions (∼ 102 − 103 times their

    original radius) and illuminate at a rate ∼ 10− 103 times higher. (Early-type stars

    cool dramatically as they expand, which compensates the increase in radius.) It is

    at this phase of evolution that stars lose substantial amounts of mass, which further

    intensifies the impact on their surroundings.

    The basic features of stellar evolution have been mostly a solved astrophysical

    problem for several decades. We are able to describe the evolving properties of

    stars quite well, by using simple conservation equations (mass and momentum), as

    well as accounting for energy generation from nuclear reactions, and radiative and

    convective energy transport (e.g. Eggleton 1971). Then the initial position of a given

    star on the Hertzsprung-Russell (HR) diagram, and its subsequent evolution, is largely

    determined by its zero age Main Sequence (ZAMS) mass (MZAMS), and metallicity

    1

  • Chapter 1. Introduction 2

    (Z).

    However, some details of the physics of stars remain poorly understood. These

    include internal rotation and its influence on stellar evolution. Stars are commonly

    assumed to rotate as solid bodies on the MS, with weak differential rotation imposed

    by convection, as is observed in the Sun. This assumption is more questionable

    post-MS. Here the specifics of the rotation profile are important, because rotation

    helps to induce mixing of chemical elements within stars, which affects the amount

    of nuclear fuel available at different stages of evolution, and increases mass loss at

    the surface by reducing the effective gravity. In addition, we have observations of

    the rotation periods of stellar remnants, which span multiple orders of magnitude in

    range (Prot ∼ 0.1− 103 days in white dwarfs and Prot ∼ 10− 103 ms in neutron stars).

    Another important piece of physics not tackled in the early stellar evolution models

    is the strength and configuration of magnetic fields. Magnetic fields are observed on

    the surfaces of stars at various evolutionary stages (from the pre-MS T-Tauri stars, to

    the MS Ap and Bp stars, to the magnetic remnants of stellar evolution, white dwarfs

    and neutron stars). Magnetic fields are not simply an observable which deserves

    an explanation, but are also able to influence the stellar evolution (primarily by

    influencing rotation, a connection which will be discussed heavily in this work). The

    combined modelling of internal stellar rotation and magnetism in evolving stars is

    still in its infancy.

    Observations continue to become more detailed, providing us with not only

    basic surface properties of stars (mass, radius, temperature, luminosity, rotation rate,

    and surface metallicity), but also information about their internal structure. One

    important observational development is the detection of propagation of waves within

    stars, obtained using asteroseismology (Unno et al. 1989; Christensen-Dalsgaard &

    Daeppen 1992). This offers fairy direct probes for the internal density, rotation, and

    energy transport (convective vs. radiative). Finally the compact remnant of a star

  • Chapter 1. Introduction 3

    offers a more indirect probe of its internal workings throughout its "life". Our goal

    here is to formulate a realistic picture of the internal rotation and magnetism of stars,

    with the minimal accuracy needed to predict the rotation and magnetic properties of

    their remnants.

    1.1 Rotation and Angular Momentum in Stars

    Up until a mere few decades ago, the only direct information regarding the angular

    momentum of stars was from observations of their surface rotation velocities (and even

    that was only the projected component - v sin i: Bernacca & Perinotto 1974, Conti &

    Ebbets 1977, and Stauffer & Hartmann 1987). Low-mass stars, with MZAMS . 1.3M�,

    lose a significant fraction of their natal angular momentum while on the MS. This is in

    contrast to more massive stars which usually retain the majority of their natal angular

    momentum. The difference arises due to the presence of a convective envelope in the

    lower mass stars, in which a dynamo is able to amplify a magnetic field. This magnetic

    field then causes particles ejected in the stellar wind to continue to co-rotate with the

    surface out to a certain distance, which effectively drains angular momentum from the

    star. This interaction between magnetic fields and ejected particles was first proposed

    by Schatzman (1962). Mestel (1968) later derived rigorous expressions to specify the

    loss of angular momentum from stars as a result of the interaction between stellar

    wind and an anchored magnetic field, and Skumanich (1972) discovered a dependence

    of surface rotation rate on age, v sin i ∝ t−1/2, known now as the Skumanich relation

    (which was calibrated in part using the solar rotation rate).

    Observations of rotation rates of evolved stars provide a weaker constraint on stellar

    models, due to the difficulty of measuring the stellar mass. An indirect measurement

    is possible in a star cluster, where the mass of the giant stars roughly corresponds

    to the mass of the MS turn-off stars. Schrijver & Pols (1993) compiled observations

    of sub-giants and giants from thirty publications. They found that stars spin-down

  • Chapter 1. Introduction 4

    much faster as they evolve off the MS, than can be expected if they conserve angular

    momentum. This effect may be partly due to the redistribution of angular momentum

    within the star, and/or to enhanced angular momentum loss to a magnetized wind.

    More precise measurements of stellar rotation became possible through astero-

    seismology (Unno et al. 1989; Christensen-Dalsgaard & Daeppen 1992). Waves

    propagating through stellar matter are influenced by internal rotation and density

    gradients, and produce detectable perturbations to the surface brightness. Therefore,

    using long cadence observations, we are able to gain a wealth of information about

    the internal properties of stars. The first star for which we obtained this insight

    was, as is usually the case, our Sun (a field specifically known as helioseismology).

    Leighton et al. (1962) were the first to definitively observe the surface perturbations

    in the Sun caused by internal acoustic waves. Later observations, made especially rich

    by the Solar and Heliospheric Observatory space telescope, provided details regarding

    the internal solar rotation (see Christensen-Dalsgaard 2002 for a detailed review of

    helioseismology and its results). We now know the Sun rotates approximately close

    to a solid body down to ∼ 0.1− 0.2R�, and particularly so in the radiative zone

    (below ∼ 0.7R�). In the convective envelope the Sun possesses a latitudinal gradient

    in rotation, with faster rotation near the equator, and slower towards the poles.

    The roughly rigid solar rotation has led many to postulate a similar rotation profile

    in all stars. And for many years this was the most reasonable choice to make, due to

    the lack of direct evidence of the internal rotation of stars. This too changed when

    instruments sensitive enough to detect the minute brightness fluctuations on the

    surfaces of distant stars became available, and especially when space telescopes with

    missions which required long cadence observations of stars were launched. CoRoT and

    Kepler have produced a wealth of information about the internal rotation of stars over

    a range of evolutionary phases. These observations clearly show that stars develop

    differential rotation as they grow towards the giant branch post-MS (Beck et al. 2012;

  • Chapter 1. Introduction 5

    Mosser et al. 2012), coinciding with the development of a deep convective envelope.

    We are therefore led to explore the possibility that deep convection zones produce

    significant radial differential rotation.

    1.2 Angular Momentum Transport

    Some of the earliest models of stellar evolution to look at internal rotation were those

    of Kippenhahn et al. (1970) and Endal & Sofia (1976). These models did not account

    for internal angular momentum transport. Starting from a solid rotation profile on the

    ZAMS, these models conserved specific angular momentum throughout. This led to

    the unphysical result that internal rotation rates surpassed the Keplerian rate (caused

    by the spin-up of the contracting core, due to Ωrot ∝ R−2 whereas ΩKep ∝ R−3/2).

    This is obviously problematic. Once angular momentum transport processes were

    incorporated into stellar models, it was possible to maintain a coupling of the core to

    the envelope, as stars expand on the giant branch (Endal & Sofia 1979).

    Angular momentum transport processes generally fit into two categories: dynamic,

    and secular. Dynamic processes occur on the rotation or convective timescales. In

    radiative layers of the star, both the magnetorotational instability (MRI - Balbus &

    Hawley 1998) and the dynamic shear instability work efficiently to dissipate shear on

    horizontal layers and enforce "shellular" rotation (as opposed to cylindrical). Secular

    processes operate on a slower timescale, comparable to the Kelvin-Helmholtz timescale

    (τKH ∼ GM2/RL) or longer. Commonly employed secular processes are meridional

    circulation and the secular shear instability, which feeds off a radial velocity gradient.

    Meridional circulation (MC), on the other hand, transports angular momentum by

    the bulk movement of matter (Maeder & Zahn 1998). MC can either occur by flow

    away from the equator along the main rotation axis, then down the surface towards

    the equator, and then back down towards the center. Or the flow can be exactly

  • Chapter 1. Introduction 6

    reverse. The former scenario transports angular momentum from the surface to the

    center, the latter scenario from the center to the surface.

    In a star, the Prandtl number is low (Pr ∼ νv/νT, with νv the viscous diffusivity

    and νT the thermal diffusivity) meaning that the thermal diffusion timescale is much

    shorter than the viscous diffusion timescale (the timescale on which angular momen-

    tum diffuses microscopically). The radiative diffusion can de-stabilize a radial angular

    velocity gradient in a stably stratified layer (Goldreich & Schubert 1967; Fricke 1968).

    In the non-axisymmetric case, this process is commonly called rotationally induced

    mixing (Zahn 1974).

    The dynamic transport processes operate on a shorter timescale, which suggests

    they dominate over the secular processes. This has made it a common practice to

    presume a "shellular" rotation profile in stellar models, an approximation we use in

    our work as well.

    The majority of models did not include the effects of magnetic torques on redis-

    tributing angular momentum, something which has recently been shown to play an

    important role. Spruit & Phinney (1998) considered it in their analysis of massive

    stars, in the context of reproducing pulsar rotation rates. They estimated the magnetic

    field strength as the minimum required to maintain solid rotation in the core, and

    posed that decoupling occurs when the rotation period becomes longer than the evo-

    lutionary timescale. Spruit (2002) formulated a more rigorous dynamo model, which

    relies on a pinch-type (Tayler) instability of an azimuthal field in a stably stratified

    region of the star. The small-scale radial field generated by the instability is then

    stretched by differential rotation. The energy lost from differential rotation pushes

    the region of the star to rigid rotation. This mechanism, commonly referred to as a

    Tayler-Spruit (TS) dynamo, has been implemented numerous times since. Heger et

    al. (2005) employed it in the Kepler stellar evolution code (Weaver et al. 1978) to

    study the effects of angular momentum transport in massive stars, with a particular

  • Chapter 1. Introduction 7

    interest on the angular momentum and magnetic fields of remnants. Cantiello et

    al. (2014) used the MESA 1-D stellar evolution code (Modules for Experiments in

    Stellar Astrophysics, Paxton et al. 2011), which includes the TS dynamo as an angular

    momentum transport mechanism. They focused on intermediate-mass stars to study

    the evolution of angular momentum as they evolve off the MS.

    Very recently, the effects of asteroseismic instabilities on angular momentum

    transport have started to be investigated. For example, Fuller et al. (2015) studied the

    effects of internal gravity waves (IGWs), and found that they are able to transport

    angular momentum quite effectively. We will not consider these effects in our work,

    as the sign of the effect varies between authors.

    1.2.1 Low- and Intermediate-Mass Stars

    Schrijver & Pols (1993) analyzed the rapid spin-down of surface rotation, observed in

    sub-giant stars as they evolve towards the giant phase. They assumed the convective

    envelope (which develops during the sub-giant phase for stars of all masses) consis-

    tently maintains a solid rotation profile, and considered rotational changes to arise

    primarily in the radiative zone below. They focused on three possible scenarios: i) the

    radiative zone rotates solidly and synchronously with the envelope; ii) it conserves

    specific angular momentum at each mass shell; or iii) it rotates rigidly but some shear

    is permitted at the convective boundary. They found that none of these three scenarios

    is able to account for the spin-down observed, and used this as evidence that strong

    magnetic wind breaking must occur to drain angular momentum from the stars. A

    similar conclusion was reached by Rutten & Pylyser (1988).

    Endal & Sofia (1979) previously considered the effect of both dynamic and secular

    transport in model stars of several masses (10, 7, 5, 3, and 1.5M�). They used the

    prescription they developed in Endal & Sofia (1976), which assumed convection zones

    maintain solid-body rotation. In this early work, they mainly focused on reproducing

  • Chapter 1. Introduction 8

    surface rotation rates, and specifically on comparing their results with two limiting

    cases: perfectly rigid rotation, and conservation of specific angular momentum by

    every shell. They found that "realistic" redistribution of internal angular momentum

    results in slower surface rotation in evolved stars than both of the above extreme

    scenarios (mainly as a result of an internal redistribution of mass by convection,

    leading to an increase in the moment of inertia). Their comparison with observed

    rotation rates of giant stars is very vague, due to poor observations available at the

    time (mere upper limits on rotation rates of giants, whose masses were unknown).

    Cantiello et al. (2014) attempted to reproduce the observations of Beck et al.

    (2012) and Mosser et al. (2012) by evolving numerical models using the MESA stellar

    evolution code. They used the built-in implementation of RIM in MESA to affect both

    angular momentum transport and chemical mixing, as well as dynamo-generated

    fields and their effect on angular momentum transport (using the TS dynamo pre-

    scription of Spruit 2002). Convection zones were treated using the mixing length

    theory (MLT), and they claimed to find that the resulting turbulent diffusivity was

    so high as to result in rigid rotation within convection zones. (They did consider a

    constant specific angular momentum profile in convection zones, but claimed it didn’t

    affect their results.) They focused their analysis on a 1.5M� star, rotating initially

    with an equatorial rate of 50km/s. They attempted several prescriptions for angular

    momentum transport (both with and without the TS dynamo, and assuming local

    conservation of specific angular momentum). However, they were unable to reproduce

    the observations of strong differential rotation of about a factor 10 between the surface

    and core of sub-giants (Beck et al. 2012), nor the dependence of core rotation period

    on stellar radius of sub-giant stars (Mosser et al. 2012).

  • Chapter 1. Introduction 9

    1.2.2 Angular Momentum Pumping by Convective Plumes and Ro-

    tational Coupling via Maxwell Stress

    We focus on two major gaps in the present understanding of angular momentum

    transport in stars. First, we consider the advection of angular momentum by radially

    extended plumes in deep convective envelopes. Second, we re-examine the role of

    Maxwell stress in redistributing angular momentum in radiative layers of a star. In

    particular, we are interested in the case where the layer is threaded by a significant

    polar magnetic flux, which is stabilized by finite magnetic helicity.

    One of the biggest modifications to internal rotation models brought on by direct

    observations, such as asteroseismology, is the ability of convection zones to generate

    differential rotation (Beck et al. 2012; Mosser et al. 2012). This is supported by fits

    of 1D stellar models to color-magnitude diagrams, and by the results of numeri-

    cal simulations of geometrically thick convection zones. Sackmann & Boothroyd

    (1991) confirmed the need for a large mixing length in giant star models. The 3D

    anelastic simulations of Brun & Palacios (2009) found that such plumes are able to

    approximately conserve angular momentum as they rise/fall.

    It is common practice to assume rigid rotation in convection zones. This is inspired

    partly by the nearly rigid rotation in the solar convection zone, and the assumption

    that the high turbulent diffusivity in convective regions is likely to efficiently smooth

    rotation gradients. We also assume that convection zones are able to transport angular

    momentum efficiently, and maintain a smooth rotation profile within them. However,

    we are guided by the numerical simulations of Brun & Palacios (2009) to allow for

    the development of differential rotation in geometrically thick convection zones (even

    reaching a constant specific angular momentum profile, Ω ∝ r−2). In geometrically

    thin convection zones we revert to a solid body rotation profile. For simplicity, we

    assume an aspect ratio Rtop/Rbottom = 2 as the boundary between thin and thick

    convection zones.

  • Chapter 1. Introduction 10

    Low- and intermediate-mass stars, those with MZAMS . 8M�, maintain a fairly

    simply structure post-MS, with a radiative core and convective envelope. We find that

    when these stars reach the tips of the red giant branch (RGB) and asymptotic giant

    branch (AGB) a constant specific angular momentum profile develops in the envelope.

    Earlier in the expansion, one must account for the backreaction of the Coriolis force

    on the inward pumping of angular momentum. When the rotation rate at the base

    of the convection zone is sufficiently fast, which we parameterize using the Coriolis

    parameter Co ≡ Ωτcon ≥ 1 (Ω is the rotation rate and τcon ≡ lP/vcon is the convective

    overturn time, defined as the time it takes a convective plume, moving at vcon, to

    traverse a pressure scaleheight lP), the Coriolis effect flattens the rotation profile.

    We derive a specific dependence of the rotation profile on the slope of gravitational

    acceleration (β, where g(r) ∝ r−β), using the "thermal-wind" approximation of the

    vorticity equation (see Section 2.2.1).

    We end up with a two-layer rotation profile in convection zones, made up of a

    slowly spinning (Co < 1) yet very steep (Ω ∝ r−2) upper part, and a rapidly spinning

    (Co ≥ 1) yet shallower lower part. We label the boundary radius between these two

    zones Rc. The rotation profile, describing the mean rotation rate over equipotential

    shells Ω̄, in the lower part of the envelope (between the base of the convective envelope

    Rbenv and Rc) is given by

    Ω̄(r) = Ω̄(Rc)(

    rRc

    )−(1+β)/2(1.1)

    =1

    τcon(Rc)

    (r

    Rc

    )−α; Rbenv < r < Rc.

    In the radiative part we find that the presence of a relatively weak magnetic field

    (with magnetic energy density ∼ 10−16 − 10−15 of the convective energy density on

    the MS, with the limit rising to ∼ 10−8 − 10−6 at later stages of evolution) is enough

    to enforce solid rotation and rotational coupling with the convective envelope. We

    therefore obtain a simple rotation profile for post-MS stars, with solid rotation in

  • Chapter 1. Introduction 11

    the radiative core, rotational coupling across the mantle, and the above discussed

    two-layer rotation profile in the convective envelope. Due to the dependence of

    rotational coupling on magnetic flux, we estimate that below a critical Co, radial

    differential rotation develops in the mantle. This is of particular interest in the rapidly

    evolving late AGB phase. In our massive star models we develop a more direct

    criterion for enforcement of solid rotation in radiative zones. In all of our models, we

    consider angular momentum transport in post-processing, thereby ignoring the effect

    of rotation on the stellar structure.

    In Section 2.3 we compare the results of our rotation model with observations

    of the core rotation rates in sub-giant stars, with a specific focus on a 1.5M� model

    rotating at 50km/s on the ZAMS (evolved using MESA and chosen to match the model

    analyzed by Cantiello et al. 2014). Such a star experiences weak angular momentum

    loss on the MS, and we assume that the same is true after it crosses the Hertzsprung

    gap. We find that our model reproduces the rotational difference between surface

    and core observed using Kepler (Beck et al. 2012), as well as the evolution of core

    rotation rates with stellar radius (Mosser et al. 2012). This comparison demonstrates

    that the best fit is obtained for α = 1, which matches our model well, because at

    this early stage of evolution, when the core is still relatively large, the dependence

    of gravitational acceleration on radius gives β ∼ 1, and hence α ∼ 1. In addition,

    we are able to reproduce the core rotation rates of helium clump stars (stars which

    are stably burning helium), and are able to tackle the drastic spin-down of surface

    rotation of stars as they evolve off the MS. By forgoing the dependence on a solid

    rotation profile, as was done by Schrijver & Pols (1993), in favour of what we consider

    a more realistic, differential, rotation profile in the deepening convective envelope we

    find the majority of the observed spin-down can be accomplished while conserving

    total angular momentum. The deepening convective envelope produces increasingly

    differential rotation, which naturally spins down the surface. However, there still

  • Chapter 1. Introduction 12

    remains a need for moderate angular momentum loss, which could occur via magnetic

    wind torques (as proposed by Schrijver & Pols 1993).

    Asteroseismology will continue to provide observational constraints on internal

    angular momentum distribution at different evolutionary stages. For example, Klion

    & Quataert (2017) considered the relative rotational splittings of both pressure and

    gravity dominated modes, which act as a probe of differential rotation in a star (Beck

    et al. 2012). They focused on sub-giant stars with masses in the range M? ∼ 1− 1.5M�,

    and radii R? ∼ 4− 5R�, and used appropriate MESA models with different rotation

    profiles applied in post-processing. They found that for many stars the rotational

    splittings are most consistent with differential rotation concentrated in the mantle

    and solid rotation in the envelope, instead of vice versa. Although, a combination of

    differential in both mantle and envelope (with α ∼ 1) is moderately consistent. We

    would like to point out three key points to consider when making a comparison of our

    rotation model and observations: i) It is plausible for strong radial differential rotation

    to exist in the mantle of our model stars. We assumed that a minimal magnetic field

    will maintain solid rotation in the entire radiative region of the star, however, there

    remains a possibility that magnetic fields could result in large-scale buoyant plumes,

    which could drive the mantle to substantial differential rotation (Nordhaus et al. 2008;

    Section 3.3); ii) The stars considered are sub-giants in which the Coriolis parameter is

    substantially greater than unity. These conditions are ripe for the development of MRI,

    which would substantially modify the rotational profile. We did not account for this

    effect in our models; and iii) So far there exists little to no observational constraints of

    the internal rotation of highly expanded giant stars. However, this is the phase along

    the giant branch that our model should be most consistent, given Co < 1 in most of

    the envelope, and the phase that is of most interest to our dynamo model (due to the

    deep penetration of the convective envelope).

    Chapter 2 discusses our rotation model in detail. In addition to the 1.5M� model

  • Chapter 1. Introduction 13

    discussed above, we consider in more detail a solar-type star (given the solar angular

    momentum when it reaches a radius R? = 10.9R� as a sub-giant), and a 5M� model

    (set to rotate at 50 km/s on the ZAMS) to test the complete evolution up to ejection of

    the envelope and emergence of the white dwarf (WD) remnant. These two masses

    were chosen to test the effects of our model in a star which contains a convective

    envelope on the MS, as well as one which contains a convective core.

    It is important to note here the codependence that exists between rotation and

    magnetic fields, each affecting the other, and both important to account for.

    1.2.3 Massive Stars

    Today there are very few direct measurements of internal rotation of massive stars,

    especially rapidly rotating ones (Pápics et al. 2012). Some indirect constraints on

    angular momentum transport come from the observable properties of their remnants,

    especially neutron stars (NSs). We will introduce the relevant properties of NSs in

    Section 1.4.2. For the immediate discussion it is only crucial to note that NSs without

    close binary companions have spin periods in the range P ∼ 10− 1000ms (Popov &

    Turolla 2012), and magnetic fields Br ∼ 3× 1011− 3× 1013G (Bhattacharya & van den

    Heuvel 1991). There are other indirect tests of internal angular momentum transport,

    such as mixing of chemical elements induced by rotation, which affects the amount

    of nuclear fuel available for burning and can contaminate the surface with heavy

    elements (Meynet & Maeder 2000), but we will not discuss these observations here.

    A comprehensive analysis of angular transport within massive stars is given by

    Heger et al. (2000). They considered several non-magnetic angular momentum

    transport mechanisms, such as dynamic and secular shear instabilities, and convective

    transport (which they assumed forces rigid rotation). They evolved stellar models of

    masses 8− 25M� from the ZAMS up to shortly before core collapse. The predicted

    NS rotation periods were significantly shorter than is observed, P ∼ 1ms in a 20M�

  • Chapter 1. Introduction 14

    model.

    This result may be due to an underestimate of magnetic torques within radiative

    zones, which would cause a net outward transport of angular momentum from the

    core. Spruit & Phinney (1998) considered a very simple model for estimating the

    magnetic flux threading the material in an evolving massive star, as was mentioned

    at the beginning of Section 1.2. They obtained extremely slow rotation in their

    remnants, P ∼ 100s, which caused them to propose supernova kicks as a prevalent

    source of remnant angular momentum. On the other hand, Spruit & Phinney (1998)

    also assumed rigid rotation in convection zones, thereby neglecting inward angular

    momentum pumping by convective plumes.

    Heger et al. (2005) incorporated a self consistent model for magnetic field

    amplification and angular momentum transport, combined with a variety of non-

    magnetic instabilities, into massive stellar models. Their dynamo prescription was

    based on Spruit (2002). They found that the magnetic field successfully enforced

    solid rotation in the inner region of their stellar models, and drastically reduced

    the angular momentum stored there (by as much as a factor of ∼ 30-50 compared

    to non-magnetic models). The relatively weak poloidal fields they obtained were

    unable to maintain core-envelope coupling beyond carbon ignition, resulting in a

    fast spinning NS (P ∼ 10ms). In contrast, Spruit & Phinney (1998) assumed perfect

    core-envelope coupling and hence obtained very slow rotation.

    Fuller et al. (2015) studied the angular momentum transport by internal gravity

    waves (IGWs). They found that IGWs can efficiently spin-down the core, relative to

    conservation of angular momentum, and obtained pulsar rotation periods & 3ms. On

    the other hand, if the core experienced significant spin-down, say due to the presence

    of magnetic torques, IGWs are likely to spin-up the core and produce an upper limit

    on the rotation period of pulsars ∼ 500ms.

  • Chapter 1. Introduction 15

    1.2.4 Angular Momentum Transport by Convection and Maxwell

    Stresses in High-Mass Stars

    Massive stars have far more complicated structures than do low- and intermediate-

    mass stars. Post-MS low- and intermediate-mass stars have a very simple structure,

    containing a convective envelope and a radiative core, surrounded by one or two

    burning shells. In contrast, massive stars posses multiple burning shells, with a

    relatively thick convection zone within each one. In fact, there could be multiple

    convection zones capable of effectively transporting angular momentum within a

    single shell. In addition, the later stages of nuclear burning are so brief that we

    must compare their duration with the timescale for angular momentum transport by

    magnetic stress. We are therefore forced to consider a more detailed version of our

    angular momentum transport mechanism.

    In our massive stellar models, the angular momentum transport timescale is

    computed within every individual mass shell. This is then compared with the

    evolutionary timescale, τev ≡ min[lP/|vr|, (tcc− t)/3] (where vr is the Eulerian velocity

    of the mass shell, and the second condition is to account for the time remaining before

    core collapse), to determine whether angular momentum is frozen or transported

    across this shell.

    Transport via convection occurs on the timescale τcon ≡ lP/vcon, where vcon is

    the convective speed. Within transporting convection zones (where τcon < τev) a

    similar rotation profile to what we introduced in Section 1.2.2 is used, with a slight

    modification. We change the strict separation between a rapidly rotating inner region

    and a slowly rotating outer region, and instead base the choice of slope of rotation

    profile on the specific value of the Coriolis parameter in each shell. The resulting

    rotation profile is

    Ω(r) ∝

    r−2 : Co(r) ≤ 1

    r−α = r−(1+β)/2 : Co(r) > 1(1.2)

  • Chapter 1. Introduction 16

    In radiative zones, angular momentum is transported via Maxwell stresses, due to an

    embedded magnetic field. Nominally, this occurs on the Alfvén timescale, given by

    τA,r = lP/vA,r, where vA,r ≡ Br/√

    4πρ is the standard Alfvén speed. However, above

    a threshold magnetic field, the hydromagnetic kink instability suppresses the rate at

    which magnetic waves are able to propagate. In Section 4.3.3 we describe the angular

    momentum transport while the kink instability grows, and in Appendix 4.8.3 we

    provide a detailed derivation. Effectively, the kink instability increases the minimum

    magnetic field required to transport angular momentum. Expressing the condition

    using the Alfvén speed, the kink instability raises the minimum speed required to

    transport angular momentum to

    vA,r

    ∣∣∣∣min

    = max

    [lPτev

    (r

    2lPΩτev

    )1/4,

    lPτev

    ], (1.3)

    where the second condition applies deep within the star in the late stages of evolution,

    when the evolutionary time is shorter than the rotation time (τev . Ω−1).

    We estimate the magnetic flux threading the material via a self-consistent model

    which leaves behind helicity at a receding convective boundary, as described in Section

    1.4.2.

    To apply our angular momentum transport model to massive stars, we consider

    four models (simulated using MESA). 13M� and 25M� models are taken as NS

    progenitors, and two 40M� models, one with solar metallicity and one with 30% solar

    metallicity, as black hole (BH) progenitors. The remnant of each model is determined

    based on the core compactness shortly before core collapse (O’Connor & Ott 2011;

    Ertl et al. 2016). We find two important conditions that determine the overall angular

    momentum distribution and evolution in our models. One is whether the star expands

    to a red or blue supergiant. The main difference between these two states is their

    surface temperature. The lower temperature in red supergiants results in a thicker

    hydrogen convection zone. This makes a significant difference to the fraction of

    angular momentum stored in the surface layers, due to increased inward pumping,

  • Chapter 1. Introduction 17

    and as a result reduces the angular momentum lost to the stellar wind. We find

    that out of our four models, the 13M� and 25M� models become red supergiants,

    whereas the two higher mass models remain blue supergiants, probably due to the

    greater mass loss (in general it remains unclear what physical properties determine

    the surface temperature of a supergiant: see Woosley & Weaver 1995; Ekström et al.

    2012). This has a significant influence on the total angular momentum available to a

    NS or BH remnant.

    The second difference between models is the quantity and thickness of convection

    zones within the burning shells of the star. A large number of thick convection

    zones implies stronger inward angular momentum pumping. We generally find

    that higher luminosities (roughly correlated with total mass) result in more effective

    angular momentum pumping. Therefore, although the total angular momentum

    of our 13M� and 25M� models is roughly the same right before core collapse, the

    angular momentum in the inner ∼ 2M� is much higher in the more massive of the

    two models.

    In Section 1.4.2 we will introduce our analysis of NS remnant rotation, as well as

    magnetic fields. And in Section 1.4.3 we will introduce our analysis of BH spins and a

    rotationally imposed cutoff to the mass collapsing to form a BH.

    1.3 Binary Interactions

    It is important to consider binary interactions when studying the evolving properties

    of stars, because most stars form in multiple systems (with the majority being binaries).

    The companions can be stellar, sub-stellar, or planetary in mass. Labelling the more

    massive of the two objects the primary, and the other one the secondary (or simply

    the companion), there are a large variety of interactions that can occur between a

    primary and secondary, depending on the physical properties of each object and the

  • Chapter 1. Introduction 18

    initial orbital configuration. The damping of tides is typically the dominant force-

    at-a-distance, and therefore tends to initiate binary interactions. These interactions

    can be mild, for example involving the transfer of angular momentum to (from) the

    orbit from (to) the rotation of either one or both companions. In the other extreme,

    the interactions could involve mass transfer, the formation of a common envelope

    surrounding both companions, and/or the eventual merger of the cores of both objects.

    Here we are primarily interested in the transfer of angular momentum from the

    binary orbit to the primary. Even seemingly negligible low-mass planetary companions

    can make a major contribution to the rotation of the primary, due to the relatively

    high specific orbital angular momentum.

    1.3.1 Low- and Intermediate-Mass Primaries

    Recent observational studies have found that ∼ 46% of solar-type stars are primaries in

    a multiple system, mostly binaries, with either a stellar or sub-stellar companion (see

    Duquennoy & Mayor 1991; Raghavan et al. 2010). However, the binary interaction

    with such massive companions could result in significant physical changes in the

    primary. We wish to avoid complicated scenarios and therefore focus on planetary

    companions to our solar-type model.

    The space telescope Kepler, whose main purpose was the detection of extrasolar

    planets, has found a plethora of planets (3,498 confirmed planets and 9,564 Objects

    of Interest as of July 20171). Focusing on planets around solar-type stars, it is found

    that ∼ 15% of such stars harbour Earth-like planets (radius ∼ 1− 1.5R⊕, see Petigura

    et al. 2013; Silburt et al. 2015; and the most recent analysis done by Fulton et al.

    2017), with orbital periods ranging from Porb ∼ 20− 200 days (the distribution of

    planet occurrences appears to be relatively flat with orbital periods). Extrapolated to

    longer periods, the prevalence seems to be ∼ 5% for Porb ∼ 200− 400 days. Larger

    1See the NASA Exoplanet Archive - http://exoplanetarchive.ipac.caltech.edu

  • Chapter 1. Introduction 19

    planets appear to be rarer, with a prevalence of ∼ 2% for planets with radii ∼ 4− 6R⊕

    (which we might consider as Neptune-like), and a prevalence of ∼ 0.14% is observed

    for planets with radii ∼ 9− 11R⊕ (which we might consider as Jupiter-like). These

    numbers for the larger planets are estimated for orbital periods shorter than ∼ 100

    days, but extrapolated to ∼ 300− 400 days they might only change by a factor of a

    few.

    Planets orbiting MS stars may be on stable, practically unchanging, orbits through-

    out the entire MS evolution. However, as the host star grows on the giant branch, the

    tidal interaction drastically increases in strength and may significantly perturb the

    orbit of the planet. It is during this phase that planets are most likely to get engulfed

    by their host stars. This phenomenon was investigated in several contexts. Kunitomo

    et al. (2011) followed up on work by Sato et al. (2008) and Villaver & Livio (2009) to

    investigate the dearth of planets within 0.6AU of clump stars, even though plenty of

    such planets exist around MS stars. They proposed that tidal interactions between

    the growing giant star and companion planets result in the engulfment of the closest

    planets. Peterson et al. (1983) considered the engulfment of planets to explain rapid

    rotation of clump stars. Livio & Soker (2002) studied the effect of the added angular

    momentum to the mass loss and magnetism of the host star. And Privitera (2016)

    considered whether planet interactions can explain the fast rotation observed in the

    surfaces of some red giant stars.

    We have considered the injection of angular momentum from a planet in combi-

    nation with the rotation model detailed in Section 1.2.2. Low-mass stars (MZAMS .

    1.3M�) lose a significant fraction of their natal angular momentum to magnetic winds

    on the MS. Therefore, they are unable to sustain the necessary spin required for an

    envelope dynamo while on the giant branch, even with the strong inward pumping of

    angular momentum in the convective envelope. These stars require an injection of

    angular momentum, which can easily be supplied by a planetary companion.

  • Chapter 1. Introduction 20

    In addition to tidal dissipation on the orbit of a planet (Hut 1981), we consider a

    gravitational perturbation induced by the time dependent density fluctuations in the

    convective envelope. We envisage that large scale convective plumes result in a time

    dependent external gravitational potential, dominated by the quadruple component.

    The magnitude of the quadrupole moment is estimated by integrating the density per-

    turbation caused by convection over the entire convective envelope. We then calculate

    the external gravitational potential of the star by randomly rotating and modifying

    the magnitude of the quadrupole moment once every global convective turnover time

    (τcon ∼ (MenvR2?/L?)1/3). Further details are given in Section 2.5. Simply put, tides

    tend to circularize and shrink the orbit, whereas the varying quadrupole stochastically

    perturbs the orbit and tends, on average, to increase its eccentricity. Therefore, the

    main effect of this varying quadrupole is to increase the range of semimajor axes over

    which tides are able to drag a companion in. See Section 2.5 for more details.

    To investigate the effect of planetary companions on a solar-type primary, we

    consider a range of planetary masses (Earth, Neptune, and Jupiter), as well as initial

    semimajor axes and eccentricities. We find that incorporating the quadrupole forcing

    increases the probability of engulfment, especially for Earth-like planets.

    We consider a simpler companion interaction in the case of the 5M� model. This

    model reaches a radius ∼ 10 times larger on the AGB, than on the RGB, which

    means that is when engulfment is most likely to take place. Therefore, to avoid

    evolving orbits of planets throughout the entire post-MS evolution, we estimate the

    tidal timescale which would result in an engulfment on the AGB. We consider the

    engulfment of planetary and sub-stellar companions with several different masses

    M ∼ 3− 100MJupiter. Even though this model does not require injection of angular

    momentum for a dynamo to be active, the increased angular momentum positively

    affects the magnetic flux amplified by the dynamo, as well as the rotation rate of the

    WD remnant.

  • Chapter 1. Introduction 21

    1.3.2 Massive Primaries

    Massive stars, the progenitors of NSs and BHs, form predominantly in multiple

    systems. Kobulnicky & Fryer (2007) found that the fraction of massive stars with com-

    panions is greater than 80%, suggesting that binary interactions are nearly ubiquitous

    (Sana et al. 2012). Much work has been done on the binary interactions of massive

    stars, typically focusing on strong interactions and end-of-life properties (core collapse,

    remnants, etc.). Binary interactions have been invoked to explain the existence of

    Type Ib/c supernovae, thought to be caused by exploding helium stars, exposed due

    to the loss of the hydrogen envelope through mass transfer onto a companion (see

    Podsiadlowski et al. 1992; Yoon et al. 2010). Binaries were also evolved beyond first

    core collapse to study the formation of double compact-star binaries. These systems

    must form after common-envelope interactions (either single or double) which cause

    the shrinking of the orbital separation between the two massive cores (see Portegies

    Zwart & Yungelson 1998; Belczynski et al. 2002). They are of interest for their

    eventual coalescence via the emission of gravitational waves (Belczynski et al. 2002;

    Abadie et al. 2010).

    In order to avoid such complicated interactions, we focus on widely separated

    binaries, which only interact via tides and never experience Roche lobe overflow. This

    process, although likely to have little effect on the stellar structure of the primary star,

    may still significantly augment its total angular momentum and affect the rotation of

    its remnant. To test the effect of this tidal interaction, we consider a simple scenario.

    For each massive model, we take a companion with half its mass, and place it in

    circular orbits of varying sizes. We evolve the orbit using the orbit-averaged tidal force

    (Hut 1981).

    We evolve the orbit concurrently with the rotational evolution of the primary,

    assuming that total angular momentum is conserved. The secondary star is assumed

    to remain unevolved, and is treated as a point object. In Sections 1.4.2 and 1.4.3 we

  • Chapter 1. Introduction 22

    discuss the binary interactions of our massive star models in more detail and present

    the result of such an interaction on the pre-core-collapse angular momentum of the

    primary.

    1.4 Rotation and Magnetism of Stellar Remnants

    After the complete exhaustion of nuclear fuel, all stars leave behind a compact remnant.

    Low- and intermediate-mass stars, those with MZAMS . 8M�, leave behind a WD

    after the ejection of their outer envelope during double shell burning. High-mass

    stars, those with MZAMS & 8M�, leave behind either a NS or a BH after their cores

    experience loss of pressure support and collapse. The type of remnant left behind

    is determined primarily by the compactness of the core shortly before core collapse

    (O’Connor & Ott 2011; Ertl et al. 2016). Below a threshold compactness, neutron

    degeneracy pressure and neutrino heating, which ramp up as the core reaches nuclear

    densities, stall the collapse and send a shockwave which causes the ejection of the

    outer layers. This results in a supernova (SN) and leaves behind a NS. At high enough

    compactness, the inner layers collapse directly into a BH. Whether this produces an

    observable electromagnetic signal depends on the angular momentum of the infalling

    material. At low angular momentum the entire star collapses into the event horizon

    (Kochanek et al. 2008; O’Connor & Ott 2011), whereas at high angular momentum a

    part of the star may become centrifugally supported, producing a relativistic outflow

    that blows up the remainder of the star (Woosley 1993).

    Even though the formation mechanism of stellar remnants is fairly well understood,

    as well as some of their basic properties (such as the equation of state of WDs), there is

    still much that is unclear. For example, the origin of rotation and magnetism observed

    in these remnants is unknown. We will now discuss each individual remnant type,

    summarize the previous work done on their rotation and magnetism, and introduce

    our model.

  • Chapter 1. Introduction 23

    1.4.1 White Dwarfs

    WDs are responsible for a variety of interesting phenomena. They are the central

    sources of illumination in planetary nebulae, are involved in binary interactions

    which create cataclysmic variables, explode as type Ia SNe (which are considered a

    standardizable candle for cosmological distance measurements, Branch & Tammann

    1992), and more. WDs are fairly well understood, due in part to their prevalence

    (Holberg et al. 2016 found 232 WDs within 25pc of the Solar System, in a sample

    which they estimated to be 68% complete), and in part to their relatively simple

    equation of state. We also understand the mechanism of formation of WDs. They

    are the inert remnant cores of stars with MZAMS . 8M�, which are exposed after the

    progenitors reach the phase of double-shell burning and eject their outer envelopes

    (producing an object observed as a planetary nebula).

    However, there are still some mysteries regarding the properties of WDs, specif-

    ically concerning their rotation and magnetism. A significant fraction of WDs, up

    to ∼ 10% by some estimates (Liebert et al. 2003; Kepler et al. 2013), are found to

    possess strong magnetic fields (B ∼ 106 − 109G). Rotation of such strongly magnetic

    WDs is observed within a wide range, from as short as one hour to as long as 10

    years (Ferrario & Wickramasinghe 2005; Brinkworth et al. 2013), with some evidence

    to suggest a positive correlation between the strength of the magnetic field and the

    rotation period (Ferrario & Wickramasinghe 2005). The origin of these fields has

    been debated for decades, with several hypotheses proposed. Fossil fields and binary

    interactions are the main two.

    In the fossil field hypothesis it is proposed that magnetic MS stars, specifically Ap

    and Bp stars, conserve magnetic flux as they evolve all the way through to envelope

    ejection (Angel et al. 1981; Ferrario & Wickramasinghe 2005). Ap and Bp stars have

    surface magnetic fields B ∼ 102 − 104G, which translates to a polar magnetic flux

    Φ ∼ B · πR2 ∼ 4× 1025− 1027Mx. This is approximately within the range observed in

  • Chapter 1. Introduction 24

    magnetic WDs. Therefore, if the magnetic flux observed at the surfaces of these stars

    also threads their cores, and it is approximately conserved throughout the evolution of

    the progenitor, strongly magnetic WDs would be a natural product. This hypothesis is

    further supported by the observation that magnetic WDs tend to be more massive than

    their non-magnetic counterparts (Kepler et al. 2013), which suggests their progenitors

    are on average more massive than those of non-magnetic WDs.

    However, strong evidence exists that magnetic WDs form exclusively as a result

    of binary interactions (Tout et al. 2010). Magnetic WDs are observed to either be

    completely isolated, or in a binary system with a close companion. In the former

    group the binary interaction could have involved the disappearance of the companion

    by a merger. In the latter group there is strong evidence that a binary interaction has

    already taken place. On the other hand, few systems have been discovered containing

    a magnetic WD with a binary companion far enough away to conclusively rule out

    any past physical interaction.

    Tout et al. (2008) suggested that progenitor systems of magnetic WDs undergo

    a common envelope interaction once the primary grows on the giant branch, which

    causes the in-spiral of the companion (either a low mass MS star or a substellar object).

    The injection of angular momentum induces differential rotation in the convective

    envelope of the giant, whence the combination of differential rotation and convection

    results in a dynamo in the envelope. In this model the strength of the magnetic field

    depends on the amount of angular momentum transferred, meaning that smaller final

    separations after the ejection of the envelope result in stronger fields as well as shorter

    rotation periods. A similar mechanism was proposed by Nordhaus et al. (2011),

    wherein the secondary (be it a planet, a Brown Dwarf or a low-mass MS star) spirals

    in close to the core where it gets shredded by the strong tidal force. This material then

    forms a Keplerian disk around the core in which a magnetic field is generated by the

    magnetorotational instability.

  • Chapter 1. Introduction 25

    We outline the issues with the above scenarios in Section 3.7.3, but briefly: in the

    former scenario convection is likely to smooth out the enhanced differential rotation

    on a fairly short timescale, and in the latter scenario a question arises about how a

    magnetic field amplified in such a transient manner can anchor in the CO core.

    Helicity Accumulation Throughout Post-MS Evolution

    We propose a mechanism by which a long term dynamo, operating at the boundary

    between the convective envelope and radiative core, can source a stable magnetic

    field in the core of a post-MS star. Previous treatments of such a dynamo have not

    accounted for long term helicity accumulation (Blackman et al. 2001; Nordhaus et

    al. 2008). The continuous mass flow across the lower boundary of the convective

    envelope, caused by single or double shell burning, gradually delivers magnetized

    material to the core. This involves a net transfer of magnetic helicity to the radiative

    material. A related configuration is encountered in a pre-MS star, and during the

    post-MS retreat of the convective core of a massive star.

    Differential rotation is the principal driver of the helicity flux. The mass flow across

    the convective-radiative boundary is itself a source of radial differential rotation. The

    strength of the radial differential rotation is enhanced by inhomogeneous rotation

    in the convection zone. A compensating Maxwell stress is then excited to enforce

    solid rotation on the radiative side. The dynamo relies on fast rotation near the

    boundary, which we parameterize using the Coriolis parameter Co. A lower bound

    on Co comes both from mean-field dynamo theory, and observations of M-dwarfs (the

    closest MS analogues of giants), which show that magnetic flux saturates at Rosby

    number Ro ≡ 2π/Co . 0.1, corresponding to Co & 60, and declines as Co2 at lower

    values. The lower threshold value of Co we use (Co . 1) is chosen due to the different

    conditions in giants compared with dwarfs, for example the much higher radiative

    flux emerging from the core (see Section 3.3.1).

  • Chapter 1. Introduction 26

    In order to quantify the remnant magnetic field in a radiative layer, we focus on the

    topological invariant H =∫

    A · BdV. Here A is the magnetic vector potential from

    which the magnetic field B is derived. H evolves only on the long ohmic timescale,

    unless an exchange of magnetic twist is possible with the exterior of the medium.

    Magnetic helicity has been shown to be an essential ingredient for maintaining

    magnetostatic stability in a static and fluid star (Braithwaite & Spruit 2004). In that

    respect, the focus on the lower boundary of a convective envelope as a source of net

    helicity is justified, as it facilitates the necessary drainage of excess helicity by the

    connection to the surface via rapid convective motions.

    We envisage a toroidally twisted magnetic field experiencing a MHD flow towards

    the core. The growth of H is proportional to the Maxwell torque (∼ BrBφr3) acting be-

    low the convective-radiative boundary. It turns out that the dominant source of such a

    torque comes from a latitudinal gradient in the Reynolds stress, imposed by convective

    motions (as seen in the solar rotation profile and in simulations of convection under

    rapid rotation, for example by Brun & Palacios 2009). This imposes a radial offset ∆Ω

    in the rotation frequency in the tachochline layer below the convective boundary. The

    winding up of the radial magnetic field is concentrated in the tachocline.

    The Maxwell stress can be estimated by using the linear winding term in the

    induction equation, to obtain a relationship between the radial (Br) and toroidal (Bφ)

    fields

    2πPdyn

    Bφ ∼r∆Ω∆r

    Br. (1.4)

    Here Pdyn is the dynamo period, which we approximate as Ndyn∆Ω−1. The toroidal

    magnetic field is in approximate equipartition with the rotational motions at the base

    of the convective envelope,

    B2φ ∼ 4πρ(

    r∆Ω2π

    )2. (1.5)

  • Chapter 1. Introduction 27

    Therefore, the Maxwell stress works out to

    BrBφ4π∼ ρr∆r(∆Ω)

    2

    2πNdyn. (1.6)

    The time derivative of H, obtained by an integral over a sphere displaced below

    the convective boundary, is expected to be non-vanishing due to a lack of perfect

    reflection symmetry about the rotational equator. There is no reason to expect that the

    distribution of magnetic flux in one hemisphere to exactly mirror the distribution in

    the other hemisphere; we only expect the total magnetic flux to vanish. More details

    regarding this general scheme are presented in Section 3.2.2.

    Another reason to focus on the stably stratified tachocline layer below the convec-

    tive envelope, is the ability of a magnetic field to anchor to this material. Magnetic

    flux tubes then rise buoyantly to retain contact with the dynamo active region, closing

    the feedback loop between toroidal and poloidal fields. The buoyancy of a flux tube is

    greatly enhanced by the strong radiative flux in an AGB star (or a massive star). It

    can be estimated from the temperature deficit induced in it by the added magnetic

    pressure (which supplements the thermal pressure). The resulting buoyant speed is

    vBr ∼(

    B2φ8πP

    )Frad

    (∆r)2ρN2. (1.7)

    P is the total pressure, ∆r is the radial distance from the convective-radiative boundary,

    and N is the Brunt-Väisälä frequency. We find that vBr is significantly faster than

    the radial speed at which mass shells sink due to nuclear processing vnucr over an

    appreciable layer below the convective boundary. This buoyancy has the effect of

    extending the active dynamo region and widening the layer with a strong differential

    rotation, the tachocline, which remains in contact with the convection zone. In Chapter

    3 we present a detailed estimate of seed field required to sustain this effect.

    We now compare two types of differential rotation as sources of the Maxwell stress

    near the convective-radiative boundary. In addition to the latitudinal gradient at the

    base of the convective envelope, driven by convection, there is the radial inflow itself,

  • Chapter 1. Introduction 28

    which drives differential rotation by the tendency of mass shells to conserve specific

    angular momentum as they sink. We label the former mechanism "Ω(θ)", and the

    latter "inflow". For both of these we estimate the growth of helicity as follows

    dHdt Ω(θ)

    = f (∆Mcore) · πεBṀcoreRbenv(ΩbenvRbenv)2, (1.8)

    and

    dHdt inflow

    = f (∆Mcore) · πṀcoreΩbenvR2benvδvr(Rbenv). (1.9)

    εB ∼ 10−3 represents the normalization of the Maxwell stress (1.6), Ṁcore is the rate

    of mass growth of the core, and δvr(Rbenv) is the speed of mass shells relative to

    the convective-radiative boundary. The factor f (∆Mcore) is meant to account for the

    cancellation effect induced by the mixing of helicity of opposite sign from the two

    hemispheres. One observes that the "Ω(θ)" contribution dominates.

    In Section 3.4.2 we present the resulting magnetic flux distribution in the cores of

    our two stellar models (solar-type and 5M�). The polar WD magnetic field lies in

    the range of BWD ∼ 106 − 108G, with the magnitude increasing proportionally to the

    injected angular momentum. The 5M� progenitor does not require an injection of

    angular momentum in order to sustain a dynamo and produce a strongly magnetized

    WD, but tidal interactions do affect the strength of the embedded field.

    We find that we are not able to reproduce magnetic fields higher than∼ 108G. In the

    solar-type star, an engulfment of a companion much more massive than Jupiter would

    require exceptional circumstances; therefore we can consider the fields obtained as

    approximate upper limits. In the 5M� star, an engulfment of a companion as massive

    as ∼ 0.1M� may even be more common, therefore fields as high as 108G can be

    expected. The extremely magnetic WDs (with fields ∼ 108 − 109G) may form from

    the merger of two WDs (García-Berro et al. 2012).

  • Chapter 1. Introduction 29

    Magnetic Field Ohmic Diffusion

    The dynamo process described so far gradually deposits magnetized material into a

    shell of limited mass forming the outer boundary of the growing core (destined to

    become a WD). In the solar-type model this occurs during both the RGB and AGB,

    when the convective envelope penetrates deep enough so as to affect layers destined to

    end up in the WD. In the 5M� model only the AGB contributes to the magnetic field

    embedded in the WD, due the shallow depth of the base of the convective envelope

    during the RGB.

    Towards the end of the evolution, as the star loses its envelope, it is also likely to

    lose enough angular momentum such that Co at the base of the envelope (Cobenv)

    drops below a critical value needed to sustain the dynamo. This means that a thin

    non-magnetic layer will prevent the magnetic field from being detectable immediately

    after the ejection of the envelope. The buried field will emerge by ohmic diffusion.

    We generalize previous calculations of ohmic decay of magnetic fields in WDs

    (Cumming 2002) to address the case of a magnetized outer shell. We find that the

    magnetic field located at the top of the magnetized layer will emerge in ∼ 107 years,

    in both mass models, suggesting a modest lag between the appearance of the WD and

    the emergence of the field. The transport time at base of the magnetized layer is ∼ 109

    years in the lower mass remnant, compared with ∼ 2× 108 years in the higher mass

    remnant. At later times, the field will decay approximately as (t/tdiff)−1/2, where tdiff

    is the diffusion timescale.

    These results suggest that in low mass WDs magnetic fields are likely to remain

    detectable until a very old age. On the other hand, the remnants of nearly all

    intermediate-mass stars will experience some field decay at ages exceeding ∼ 109

    years.

  • Chapter 1. Introduction 30

    Rotation of White Dwarfs

    Turning now to the rotation period in the remnant WD, we emphasize that it depends

    on the angular momentum retained by the core past the phase of envelope ejection.

    As the star loses mass it also loses angular momentum, and as long as the core

    remains rotationally coupled to the envelope, the core will lose angular momentum as

    well. Given that rotational coupling requires a minimum magnetic flux, the strong

    dependence of the dynamo-generated field on Cobenv (as suggested by observations

    of MS fields, e.g. Reiners et al. 2009) means that that when Cobenv drops below a

    critical value, the magnetic flux amplified by a dynamo becomes too weak and the core

    decouples. The resulting rotation period depends on the choice of this critical value;

    we envision that decoupling should be possible for Cobenv < Cobenv,crit ∼ 0.1− 0.3.

    Under the assumption of sustained rotational coupling, the core angular momen-

    tum, Jcore, in the 5M� model remains roughly constant as the star loses its envelope

    at the end of the AGB, due to the deep convective envelope maintaining a constant

    specific angular momentum profile in the envelope. This means that the angular

    momentum in the remnant (which we set equal to Jcore) is approximately independent

    of when the core decouples. Therefore, the remnant spin is directly proportional

    to the initial angular momentum of the star (except in the case of significant angu-

    lar momentum injection, such as by the absorption of a massive companion, say

    & 3MJupiter).

    An interesting change in rotation results from a re-expansion of the envelope

    caused by a late helium flash, which is seen in our solar-mass model. During this

    flash ∼ 10−3M� of radiative material expands into a large convective envelope, which

    results in a final loss of mass, and a significant spin-down of the star. During the main

    AGB wind phase the stellar angular momentum drops by a factor ∼ 103, whereas the

    core angular momentum drops by a factor ∼ 5. But during the late re-expansion, both

    stellar and c