The actions of non-equilibrium systems and related matters

Michael Joseph Landry

Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy under the Executive Committee of the Graduate School of Arts and Sciences

COLUMBIA UNIVERSITY

2021 © 2021

Michael Joseph Landry

All Rights Reserved Abstract

The actions of non-equilibrium systems and related matters Michael Joseph Landry

In this work, we develop an effective field theory program for many-body systems out of finite temperature equilibrium. Building on recent work, we combine powerful mathematical tools such as the Schwinger-Keldysh closed-time-path formalism, the coset construction, and Wilsonian effective field theory to construct novel actions that describe a wide range of many- body systems out of finite-temperature equilibrium. Unlike ordinary actions, these non- equilibrium actions account for dissipation and statistical and quantum fluctuations. The novel actions constructed include those for , , nematic crystals, smectic liquid crystals in phases A, B, and C, chemically reacting fluids, quasicrystals, higher-form dual theories of superfluids and solids, and plasmas that can support large charge density. In order to construct these actions, we propose a new kind of coset construction with a total of four distinct types of inverse Higgs constraints. We extend the coset construction to account for higher-form symmetries and investigate the relationship between two kinds of ’t Hooft anomalies and spontaneous symmetry breaking. Table of Contents

Acknowledgments ...... v

Dedication ...... vii

Chapter 1: Introduction and Background ...... 1

1.1 What is a non-equilibrium system? ...... 1

1.2 Why bother with effective field theory? ...... 3

1.3 Zoology and effective field theory ...... 5

1.4 Non-equilibrium effective actions ...... 8

1.4.1 KMS conditions and thermal equilibrium ...... 11

1.4.2 Rules for constructing non-equilibrium EFTs ...... 12

1.5 Diffusion action ...... 15

1.6 Relativistic hydrodynamics ...... 19

1.6.1 The standard approach ...... 19

1.6.2 The fluid worldvolume ...... 22

1.6.3 The action approach ...... 24

1.7 The meaning of retarded and advanced fields ...... 28

1.8 Organization of the thesis ...... 29

Chapter 2: The coset construction for non-equilibrium effective actions ...... 31

i 2.1 The zero-temperature coset construction: a review ...... 34

2.1.1 Inverse Higgs ...... 37

2.1.2 Zero-temperature superfluids: a simple example ...... 38

2.2 The non-equilibrium coset construction ...... 40

2.2.1 Goldstones at finite temperature ...... 41

2.2.2 The method of cosets ...... 42

2.3 Fluids and superfluids at finite temperature ...... 48

2.3.1 Fluids ...... 48

2.3.2 Superfluids ...... 53

2.4 Solids and supersolids at finite temperature ...... 55

2.4.1 Solids ...... 55

2.4.2 Supersolids ...... 58

2.5 Liquid crystals at finite temperature ...... 60

2.5.1 Nematic liquid crystals ...... 62

2.5.2 Smectic liquid crystals ...... 66

2.6 Conclusion ...... 71

Chapter 3: Higher-form symmetries and ’t Hooft anomalies in non-equilibrium systems 73

3.1 Introduction ...... 73

3.2 Higher-form symmetries: a review ...... 75

3.2.1 Superfluids at zero temperature ...... 78

3.2.2 Electromagnetism ...... 80

3.3 Non-equilibrium cosets and the classical limit ...... 81

ii 3.4 Non-equilibrium ’t Hooft anomalies ...... 85

3.4.1 Reactive fluids ...... 87

3.4.2 Second sound and its removal ...... 91

3.4.3 Dissipative solids and smectics without second sound ...... 94

3.4.4 Quasicrystals ...... 99

3.5 Higher-form symmetries and the coset construction ...... 105

3.5.1 The zero-temperature limit ...... 108

3.5.2 Gauge theories ...... 110

3.6 Higher-form symmetries and ’t Hooft anomalies ...... 116

3.6.1 Gauge fields and the Maurer-Cartan form ...... 116

3.6.2 Dual theories and mixed ’t Hooft anomalies ...... 119

3.6.3 Non-equilibrium higher-form ’t Hooft anomalies ...... 121

3.6.4 Quadratic examples ...... 123

3.7 Dual superfluids ...... 133

3.7.1 Zero temperature ...... 133

3.7.2 Finite temperature ...... 136

3.8 Dual solids ...... 138

3.8.1 Zero temperature ...... 138

3.8.2 Finite temperature ...... 141

3.9 Discussion ...... 144

Epilogue ...... 147

References ...... 150

iii Appendix A: Emergent gauge symmetries ...... 157

Appendix B: Stückelberg tricks and the Maurer-Cartan form ...... 159

Appendix C: Explicit inverse Higgs computations ...... 162

C.1 Thermal inverse Higgs ...... 164

C.2 Unbroken inverse Higgs ...... 164

C.3 More on unbroken inverse Higgs ...... 166

Appendix D: Fluids and volume-preserving diffeomorphisms ...... 168

iv Acknowledgements

There are may people who deserve acknowledgment and thanks both for my academic progress and my personal development. First I would like to thank my advisors, both official and unofficial. Thank you Alberto Nicolis and Lam Hui for your wonderful mentorship through out my graduate education. It has been a privilege learn from you and I have benefitted immensely from your uncommon talents. You managed to simultaneously teach me enormous amounts of physics while fos- tering an environment of independence. I truly could not have hoped for better advisors. Additionally, I thank Matteo Baggioli for the many projects and discussions we have shared. Your enthusiasm for physics is contagious and over the past year, you have broadened my un- derstanding of physics enormously. Austin Joyce deserves a big thank you from not only me but all of the other graduate students in the theory department. Your patience in explaining complicated concepts in a clear way is unparalleled. You have done so much for us and asked for so little in return. Additionally, I am extremely grateful to my undergraduate advisor, Cumrun Vafa for your enormous support, both professional and personal. Thank you to my high school physics teacher Michael Stark for your enthusiasm and encouragement during the earliest stages of my development as a physicist. In high school, I worked on a research project with Professor Dorian Goldfeld. I thank you for giving me my first introduction to research; you taught me an enormous amount and I am truly grateful for everything you have done for me. Lastly I would like to thank Professor Morgan May for encouraging my curiosity and indulging my many questions about the physical universe since I was a child.

v A big thank you to my fiends Noah Bittermann, Roman Berrens, Guanhao Sun, Benjamin Gilbert, Steve Harrelson, and Farzan Vafa for the many discussions we have had. It has been a pleasure working with all of you. I have been extremely lucky to be surrounded by a loving and supportive family. None of this would have been possible without them. Thank you to both my American family for your unconditional love and support and thank you to my Russian family for accepting me as one of your own. A special thank you to my dad for teaching me calculus when I was twelve and to my mom for teaching me to read despite my best efforts. Thank you to Chris for being the best big brother I could have hoped for. And thank you to Anastasia for accompanying me, supporting me, and loving me every step of the way. Finally, thank you to Uncle Bob for everything (there is far too much to list here). Lastly, I am deeply grateful to the many doctors, nurses and other medical professionals without whom I would not be here.

vi Dedication

To my loving wife, Anastasia

vii Chapter 1: Introduction and Background

The aim of this thesis is to investigate the non-equilibrium dynamics of many-body systems from the perspective of effective field theory. In Chapter 1, we will review some important results derived in recent works [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28]. Chapters 2 and 3 contain original work that I completed as a graduate student. These chapters are primarily based on [29, 30] but also contain summaries of important results from [31, 32, 33]. Finally, as a graduate student, I also took part in [34, 35, 36], but in the interest of brevity, I will not discuss these works.

1.1 What is a non-equilibrium system?

Perhaps a better question is: What is not a non-equilibrium system? There are two categories of physical systems that do not fall under the heading of non-equilibrium physics. The first includes equilibrium systems, most of which are in thermal equilibrium. The second includes zero-temperature systems, which are described by pure quantum states. In reality, both of these categories are idealizations; no system is ever in perfect equilibrium or exactly at zero temperature. In this sense, non-equilibrium physics includes all physical phenomena that are not subject to these idealized assumptions. The focus of this thesis will be the behavior of non-equilibrium phenomena with very many degrees of freedom. Such systems are often referred to as many-body systems. The behavior of many-body systems are often highly chaotic and complicated; they therefore resist simple analytic description. If, however, we consider only the behavior of these systems over long distance and time scales, the descriptions drastically simplify. The reason is that over sufficiently long scales, equilibrium is obtained in an approximate, local

1 sense. That is, much of the the complicated, chaotic behavior typically dies down over a time-scale known as the relaxation time. Thus, considering the behavior of systems on scales much longer than the relaxation time allows us to sweep much of the chaotic ugliness under the rug and define an effective description that involves only a small number of degrees of freedom. As a result, we may treat each (sufficiently large) local region as if it were in a state of approximate equilibrium. The focus of this thesis will be to consider such large-scale behavior, allowing us to restrict attention to near-equilibrium states. Borrowing terminology from high-energy physics, we will refer to the large-scale physics as infrared (IR) and the small-scale physics as ultraviolet (UV). Because we are interested in IR physics, we will focus on gapless (or nearly gapless) excitations, which exist at arbitrarily small frequency and wave number. It turns out that symmetry considerations are an important—if not the most important—guiding principle when constructing effective theories of IR physics. The reason symmetries play such an important role has to do with two of the most important theorems in physics: Noether’s theorem and Goldstone’s theorem. (1) Noether’s theorem states that if an action enjoys a continuous symmetry, then there exists a corresponding current that is conserved on the equations of motion. Moreover, there is a one-to-one relationship between conserved currents and continuous symmetries. In non-equilibrium systems, Noether currents play a very im- portant role because any non-uniformities in the currents are prohibited from equilibrating locally by their conservation equations. A classic example is particle diffusion in a liquid. A splotch of red die will very slowly diffuse throughout a glass of water because the number of die molecules is conserved; if these molecules were not conserved, then the die would simply decay and would not persist over long times or distances. (2) Goldstone’s theorem states that if a system spontaneously breaks a continuous symmetry, then there exists a corresponding gapless excitation known as a Goldstone boson. Spontaneous symmetry breaking (SSB) oc- curs when the equilibrium (or ground) state of a system transforms non-trivially under a given symmetry. If Poincaré symmetry is unbroken, then Goldstones’s theorem guarantees

2 a one-to-one relationship between spontaneously broken symmetries and Goldstone bosons. If Poincaré symmetry is spontaneously broken, then no such one-to-one correspondence is required, but at least one gapless Goldstone mode is guaranteed. The above approach to physics in which IR degrees of freedom are considered, while UV modes are ignored, is known as effective field theory (EFT). We will use this EFT philosophy, grounded in symmetry principles, to construct effective actions that describe the dissipative and stochastic dynamics of many-body systems out of finite-temperature equilibrium.

1.2 Why bother with effective field theory?

There are entire subfields of physics devoted to investigating the various systems that we will consider in this thesis, so some readers might wonder: What is the point of approaching these topics from an EFT perspective? There are several possible responses: Firstly, from a purely theoretical point of view, these EFT formalisms are elegant and beautiful in their paucity of assumptions. The only input for these theories is the symmetry- breaking pattern of the equilibrium state. Once this single piece of information is known, all else follows. Secondly, this lack of choice in constructing EFTs has a clear practical benefit: As long as we are careful to include all terms consistent with symmetries, we can be confident that our theory is complete (at any given order in a derivative expansion). More traditional methods may depend on phenomenological assumptions or physical intuitions. As a result, there is no guarantee that all possible terms have been accounted for. In this way, we can use the EFT formalism to check existing theories for missing terms, thereby discovering new physics. Third, by casting the theories for many-body systems in terms of an action principle, we are able to drastically simplify otherwise complicated problems. While it would be possible to describe particle physics using methods other than an action approach, few would argue for doing so as actions allow for exceedingly economical descriptions of myriad phenomena. Similarly, when employing effective actions to describe many-body systems,

3 we expect comparable calculational benefits. A particularly elegant example is the use of an action to describe the precession of a vortex line in a trapped superfluid [37]. While such a system could be understood via other means, e.g., numerical simulations, the EFT description reduces the question of how the vortex precesses to just two easily-evaluated Feynman diagrams. Lastly, if we wish to couple our EFTs to other forms of matter, the process for doing so is straight-forward; simply write down all coupling terms consistent with symmetries. A powerful application of these EFT principles involves the direct detection of dark matter. In recent works [38, 39, 40, 41], EFTs that describe the couplings of various dark matter models with condensed matter systems are constructed, thereby providing definite exper- imental predictions. The method for arriving at these experimental predictions is quite straight-forward; there is no guess-work at any stage. First, the action for the condensed matter system and its coupling to the postulated dark matter field is constructed from sym- metry principles. At any given order in the derivative expansion, there will be finitely many coefficients, which may be experimentally determined.1 Next, from this action, Feynman diagrams can be constructed. The rules for computing Feynman diagrams are well-studied and are easy to implement. Once these Feynman diagrams have been computed, scatter- ing amplitudes follow immediately. Finally, these scattering amplitudes can be converted to scattering cross sections, which describe the rate at which dark matter particles scatter off of the condensed matter system. These cross sections are observables, meaning that the theory’s predictions can be experimentally tested. It is worth noting that there are many other methods for calculating these observables; however, the EFT procedure is particularly simple. Given the SSB pattern of the condensed matter system and the kind of dark matter particle postulated by a given theory, the process by which scattering cross sections are cal- culated is reducible to turning a mathematical crank. Thus otherwise challenging problems are solved by performing almost trivial calculations. 1In an ordinary fluid, examples of such coefficients include the speed of sound, shear and bulk viscosities, and thermal conductivity.

4 While the EFT approach offers a more unified and systematic method for constructing theories, it does not necessarily yield the best tool for solving all physics problems. In particular, Feynman diagrams offer a convenient way to organize a perturbative expansion, but some phenomena may fall outside of this perturbative regime. For example, results pertaining to highly non-linear hydrodynamics may be ill-suited for a expansion. As such, more traditional analytic or numerical methods may be required. At this moment, the EFT approach to many-body systems is still quite new. Conse- quently, its full potential has yet to be realized. However, given its clear conceptual and practical advantages, it is reasonable to expect that it will continue to yield important re- sults.

1.3 Zoology and effective field theory

It is often the case that all gapless excitations of a non-equilibrium system are associated with Noether currents or are Goldstone bosons. In this thesis, we will restrict attention to such systems, though it should be noted that there are systems for which this rule does not apply. Systems at critical points of transitions, for example, can exhibit gapless excitations that are not classifiable in terms of Noether currents and Goldstones. The reasons to focus on systems for which symmetries are the only guiding principle is (1) it simplifies matters greatly, (2) the Poincaré group is always spontaneously broken by a many-body system; at a minimum, the zero-momentum frame breaks Lorentz boosts and (3) it is a valid assumption for an enormous range of systems. In fact, the SSB pattern of a system is often used to classify its . When constructing EFTs of condensed matter systems, it is often necessary to introduce additional symmetries Q. These symmetries allow, as an example, for some long-distance notion of homogeneity and isotropy even when translations and rotations are spontaneously broken. At zero temperature there are eight broad states of matter characterized by their own unique SSB pattern; we enumerate them below [1]. The Poincaré algebra in 3+1 dimensions

5 is

i[Jµν, Jρσ] = ηνρJµσ − ηµρJνσ − ησµJρν + ησν Jρµ,

i[Pµ, Jρσ] = ηµρPσ − ηµσ Pρ, (1.1)

i[Pµ, Pν] = 0,

1 ijk where Pµ generate spacetime translations, Ji ≡ 2  Jjk generate spatial rotations, and Ki =

J0i generate Lorentz boosts. We denote unbroken translations and rotation by P¯0, P¯i, and J¯i for i = 1,... 3. These unbroken translations and rotations may involve internal symmetries Q. We nevertheless must require that they satisfy

i jk i jk [J¯i, J¯j] = i J¯k, [J¯i, P¯j] = i P¯k . (1.2) if large-scale isotropy and homogeneity are to be restored. The states of matter are as follows:

• P¯0 = P0, P¯i = Pi, J¯i = Ji

This state of matter enjoys the minimal SSB pattern; only boosts are spontaneously broken, so we expect three Goldstone bosons. Following [1], we call a medium described by this SSB pattern a type-I framid. Interestingly, type-I framids are nowhere to be found in nature. The reason for their absence remains a mystery.

• P¯0 = P0 + Q, P¯i = Pi, J¯i = Ji

In the minimal case, we take Q to be the generator of a U(1) symmetry, which has the physical interpretation of particle number. This SSB pattern is associated with ordinary superfluids, which have been termed type-I superfluids.

• P¯0 = P0, P¯i = Pi + Qi, J¯i = Ji

Because the unbroken generators must satisfy (1.2), the internal generators Qi must enjoy non-trivial commutation relations with the spatial rotation generators of Poincaré

6 group. That is, Qi cannot be internal symmetry generators. Such a state of matter is known as a type-I galileid.

• P¯0 = P0, P¯i = Pi, J¯i = Ji + Q˜i

i jk The commutation relations (1.2) require that [Q˜i, Q˜ j] = i Q˜k, meaning that these

Q˜i generate an internal SO(3) group. In this scenario, we have spontaneously broken boosts and rotations. As a result, there are six Goldstone bosons. Such a state of matter is known as a type-II framid.

• P¯0 = P0 + Q, P¯i = Pi + Qi, J¯i = Ji

Once again, (1.2) require that Qi have non-trivial commutation relations with Poincaré rotations. As a result, this sate of matter bears a striking resemblance to the galileid. We therefore term is the type-II galileid.

• P¯0 = P0 + Q, P¯i = Pi, J¯i = Ji + Q˜i

As before, (1.2) implies that Q˜i generate an internal SO(3) group. As a result, we have merely added symmetries that commute with spacetime translations to the superfluid case. We therefore term this state of matter a type-II superfluid.

• P¯0 = P0, P¯i = Pi + Qi, J¯i = Ji + Q˜i

Requiring (1.2) and that Qi and Q˜i commute with all Poincaré generators, we have

ijk ijk [Q˜i, Q˜ j] = i Q˜k, [Qi, Q j] = 0, [Q˜i, Q j] = i Qk . (1.3)

These commutation relations match those of the three-dimensional Euclidean group

ISO(3). In particular, the Q˜i’s generate an internal SO(3) symmetry and the Qi’s generate three-dimensional internal translations. Such a medium can be interpreted as an isotropic [1].

• P¯0 = P0 + Q, P¯i = Pi + Qi, J¯i = Ji + Q˜i

7 This state of matter looks almost identical to that of the solid above. The one addition is the spontaneous breaking of time translations with the introduction of the U(1) generator Q. The SSB pattern is a combination of those of the superfluid and the solid. We therefore call the resulting state of matter a .

The list of states of matter itemized above is by no means exhaustive. In particular, there exist other states of matter like , liquid crystals, fermi liquids, string fluids, and many more. In this thesis we will see how effective field theories of many of these other states of matter can be constructed from SSB considerations.

1.4 Non-equilibrium effective actions

At zero temperature, the equilibrium state is described by a pure state, namely the vacuum. At finite temperature, however, no such pure equilibrium state exists. Instead, the equilibrium state must be described by a mixed-state thermal density matrix of the form

−β P¯0 e 0 ¯0 ρ = , Z = tr e−β0P , (1.4) Z

0 where β0 is the inverse equilibrium temperature and P¯ is the unbroken time-translation generator. In order to describe correlation functions in time, we must use the so-called in-in or Schwinger-Keldysh formalism [42]. In this formalism, the sources are doubled. Letting U(t, t0, J) be the time-evolution operator from time t0 to t in the presence of source J for some field Ψ, the generating functional is

W[J1,J2]  †  e ≡ tr U(+∞, −∞; J1)ρU (+∞, −∞; J2) ¹ (1.5) iS[Ψ1,J1]−iS[Ψ2,J2] ≡ DΨ1DΨ2e , ρ where S is the action of the ultraviolet theory. In the path integral representation in the above equation, we require that, in the distant future, Ψ1(t → ∞) = Ψ2(t → ∞), and the subscript ρ indicates that field configurations are weighted by the thermal density matrix

8 functional in the infinite past. Because there are two time-evolution operators such that one evolves forward in time and the other backward in time, we can conceive of the SK path integral as existing on a closed contour in time that starts at t = −∞ goes to t = +∞ and then returns again to t = −∞. This contour is often referred to as a closed-time path (CTP). It is possible to use this formalism to construct an effective action of the IR degrees of freedom on the CTP, known as the non-equilibrium effective action. Since the generating functional depends on two copies of the sources, the non-equilibrium effective action will have doubled field content. Consider the path-integral representation of the generating functional (1.5). We are interested in the meaning of the low-frequency, long-wavelength dynamics of the system. In typical Wilsonian fashion, we integrate out the fast modes to obtain an effective action for the slow modes. Suppose the fields can be divided up into IR and UV fields by Ψ = {ψir, ψuv }, where ψir represent the IR degrees of freedom and ψuv represent the UV degrees of freedom. Define the effective action for the IR fields by2

ir ir ¹ uv ir uv ir iIEFT[ψ1 ,ψ2 ;J1,J2] uv uv iS[ψ1 ,ψ1 ;J1]−iS[ψ2 ,ψ2 ;J2] e = Dψ1 Dψ2 e , (1.6) ρ

ir ir W[J1,J2] ∫ ir ir iIEFT[ψ1 ,ψ2 ;J1,J2] then e = D[ψ1 , ψ2 ]e . Notice that the information contained in the

density matrix ρ is absorbed into the coefficients of IEFT so we do not need to include a

ir density matrix for the IR fields ψ1,2. Finally we require that the Green functions of IEFT be path-ordered on the SK contour; the path-ordered Green functions are given as follows. Suppose we focus on the quadratic part of W, which we denote by W(2). Then we have that

¹ ¹ GF −iG− J (x ) (2) i d+1 d+1   © ª © 1 2 ª W [J1, J2] = d x1 d x2 J1(x1) J2(x1) ­ ® ­ ® . (1.7) 2 ­ + F ® ­ ® −iG G˜ J2(x2) « ¬ « ¬ 2Here as well as in the remainder of the thesis, we use S to denote an ordinary action and I to denote an action defined on the SK contour.

9 In the previous equation we employed the definitions

F G (x1, x2) = ihTO(x1)O(x2)i,

F G˜ (x1, x2) = i T˜O(x1)O(x2) , (1.8) + G (x1, x2) = hO(x1)O(x2)i,

− G (x1, x2) = hTO(x2)O(x1)i,

where O(x) is the operator sourced by J(x), and T and T˜ represent time-ordering and anti- time-ordering, respectively. Importantly, O need not be the fundamental quantum field Ψ; depending on how J appears in the action, it may be some complicated composite operator. The fact that the Green functions take these forms follows directly from the path-ordering prescription on the SK contour. Now suppose we work in a different basis defined by

1 Jr ≡ (J1 + J2), Ja ≡ J1 − J2. (1.9) 2

Then the quadratic part of the generating functional is somewhat simplified, namely

¹ ¹ 0 GA J (x ) (2) i d+1 d+1   © ª © r 2 ª W [Jr, Ja] = d x1 d x2 Jr (x1) Ja(x1) ­ ® ­ ®, (1.10) 2 ­ R S® ­ ® G ˜iG Ja(x2) « ¬ « ¬

where

1 GS(x , x ) = hO(x )O(x ) + O(x )O(x )i, 1 2 2 1 2 2 1

∆(x1, x2) = h[O(x1), O(x2)]i, (1.11) R G (x1, x2) = iθ(t1 − t2)∆(x1, x2),

A G (x1, x2) = −iθ(t1 − t2)∆(x1, x2).

GR is the retarded propagator, GA is the advanced propagator, and GS is known as either

10 the symmetric or Keldysh propagator.

1.4.1 KMS conditions and thermal equilibrium

We will now investigate the implications of having a thermal density matrix as the equi- librium state. We will derive the well-known KMS conditions of thermal equilibrium. Rec- ognizing H ≡ P¯0 as the Hamiltonian, we see that for any real number a, we may treat

aH −i(ia)H e = e as the time-evolution operator by amount ia. Let U1 ≡ U(J1) ≡ U(+∞, −∞; J1)

and U2 ≡ U(J2) ≡ U(+∞, −∞; J2). Then the generating functional for ρ given by (1.4) is

1 1 eW[J1,J2] = tr e−β0HU†U  = tr e−(β0−θ)HU†e(β0−θ)H e−β0H eθHU e−θH  Z 2 1 Z 2 1  †  = tr U (J2(t − i(β0 − θ)))ρU(J1(t + iθ)) (1.12)

≡ eWT [J1(t+iθ),J2(t−i(β0−θ))].

for any constant θ ∈ [0, β0]. Expanding the above equation to quadratic order in the sources gives the relation

1 β0ω GS(ω, kì) = coth ∆(ω, kì), (1.13) 2 2

which is known as the fluctuation-dissipation theorem. Unfortunately, for higher-point cor-

relation functions, no constraints on W can be deduced as WT expresses a different set of correlation functions than W.

But now suppose that the UV theory enjoys a discrete Z2 anti-unitary time-reversing symmetry Θ, that is [Θ, H] = 0. Here Θ could be simply the time-reversal operator T , but more generally, it is any combination of C and P with T . Combining Θ with (1.10), we arrive at the equality

W[J1(x), J2(x)] = W[J˜1(x), J˜2(x)], (1.14)

11 where

J˜1(x) = ΘJ1(t − iθ, xì), J˜2(x) = ΘJ2(t + i(β0 − θ), xì), (1.15) for constant θ ∈ [0, β0]. In the next subsection, we will see that these KMS conditions have an analogue in the effective actions.

1.4.2 Rules for constructing non-equilibrium EFTs

Following the usual EFT philosophy, it is our hope that all of the complicated UV dynam- ics and information about ρ can be absorbed in the low-frequency limit by a finite number of parameters in IEFT at any given order in the derivative and field expansions. It has been demonstrated that this is in fact the case [2], but there are several important constraints that must be imposed upon IEFT. We outline some essential features of this effective action below without proof [2]:

3 • The UV action describing the system of interest is factorized by S[Ψ1; J1] − S[Ψ2; J2]. The effective action, however, does not admit a factorized form. In general, there will

exist terms that couple 1-and 2-fields in IEFT.

• Notice that, while the coefficients of S[Ψ1; J1]−S[Ψ2; J2] are purely real, the coefficients

ir ir of IEFT[ψ1 , J1; ψ2 , J2] may be complex. There are three important constraints that come from unitarity, namely

∗ ir ir ir ir IEFT[ψ1 , ψ2 ; J1, J2] = −IEFT[ψ2 , ψ1 ; J2, J1]

ir ir ir ImIEFT[ψ1 , ψ2 ; J1, J2] ≥ 0, for any ψ1,2, J1,2 (1.16)

ir ir IEFT[ψ1 = ψ2 ; J1 = J2] = 0.

These conditions help to ensure that Green functions are path-ordered.

• Any symmetry of the UV action S is a symmetry of IEFT, except for time-reversing

3When we say that such an expression “factorized,” we really mean that its exponential eiS[Ψ1;J1]e−iS[Ψ2;J2], which appears in the path integral, is factorized.

12 symmetries. The fact that these time-reversing transformations are not symmetries of the effective action allows the production of entropy. Because the field values on

ir ir the 1-and 2-contours must be equal in the distant future, ψ1 and ψ2 must transform simultaneously under any global symmetry transformation. Thus, there is just one copy of the global symmetry group.

¯ • If the equilibrium density matrix ρ takes the form of a thermal matrix, ρ ∝ e−β0P0 , then

the partition function W[J1, J2] obeys KMS conditions (1.15). These KMS conditions for the partition function can be used to derive the so-called dynamical KMS symme- tries of the effective action. The way these symmetries act is as follows. Suppose that the UV theory possesses some kind of anti-unitary time-reversing symmetry Θ. Then, setting the sources to zero, the dynamical KMS symmetries act on the fields by

ir ir ψ1 (x) → Θψ1 (t − iθ, xì) (1.17) ir ir ψ2 (x) → Θψ2 (t + i(β0 − θ), xì),

for any θ ∈ [0, β0]. It can be checked that these transformations are their own inverse,

meaning that the dynamical KMS symmetries are discrete Z2 symmetries. These sym- metries involve temporal translations along the imaginary-time directions because the thermal density matrix can be interpreted as a time-translation operator by imaginary

time −iβ0. As a result, they are non-local transformations. This non-locality may seem rather odd, but at any given order in the derivative expansion, these symmetries

become local; one need only perform a Taylor series in θ and β0 − θ. Further, in the classical limit they become exactly local. To take the classical limit, it is convenient to perform a change of field basis by

1   ψir ≡ ψir + ψir , ψir ≡ ψir − ψir . (1.18) r 2 1 2 a 1 2

13 Then the classical dynamical KMS symmetry transformations are

ir ir ψr (x) → Θψr (x) (1.19) ir ir  ir  ψa (x) → Θψa (x) + iΘ β0∂tψr (x) .

ir ir Notice that the change in ψa is proportional to the derivative of ψr . Thus, when writing down terms of the effective action in the derivative expansion, it is natural to

ir ir consider ψa and ∂tψr as contributing to the same order.

Finally, applying a Noether-like procedure to the classical dynamical KMS symmetries, it is possible to construct a current sµ whose divergence on-shell is always nonnegative. This current can be identified with the local entropy current, and the nonnegative divergence enforces the local statement of the second law of thermodynamics [3]. Consider the effective ∫ 4 action without sources in the classical limit IEFT[ϕr, ϕa] = d x LEFT[ϕr, ϕa]. To keep things fully general, suppose the classical dynamical KMS transformations are

ϕr (x) → Θϕr (x) (1.20) ϕa(x) → Θϕa(x) + iΘΛr (x),

for some r-type field Λr . Then under a dynamical KMS transformation, the effective La-

µ grangian can change by at most a total derivative, LEFT → LEFT + ∂µV . We can expand V µ in powers of a-type fields µ µ µ V = iV0 + V1 + ··· , (1.21)

µ where Vk contains k factors of a-type fields. Since the dynamical KMS symmetry is discrete

and terms of LEFT must all have at least one a-type field, it is possible (using integration by

µ µ parts) to define LEFT such that V = iV0 . However, if we express LEFT in such a way that the terms linear in a-type fields have no derivative acting on the a-type fields, then we may µ µ ≥ have a non-zero V1 but still have Vk = 0 for k 2. In this case, define the entropy current

14 by µ µ ˆ µ s = V0 − V1 , (1.22)

ˆ µ µ| ≥ where V1 = V1 ϕa=Λr . Using the fact that ImIEFT 0 it has been demonstrated in [3] that µ ∂µs ≥ 0, meaning that the second law of thermodynamics is automatic.

1.5 Diffusion action

To illustrate how a non-equilibrium effective action can be constructed, we will now consider a simple example. Suppose a system in which there is a single conserved U(1) charge—this charge could correspond to conserved particle number—that exists at inverse

µ 4 temperature β0. Let J be the corresponding Noether current. Since we are assuming that nothing but the U(1) charge is conserved, we must consider a system that explicitly breaks Lorentz boosts. We allow our system to enjoy spacetime translation symmetry and spatial rotation symmetry.5 We would like to identify the collective IR degrees of freedom associated with correlations of the conserved current J µ. We therefore consider the generating functional for correlations functions of J µ insertions on the SK closed-time-path

 ∫ d µ µ  W[A1µ,A2µ] i d x[A1µ J −A2µ J ] e = Tr ρ0Pe 1 2 , (1.23)

where ρ0 denotes the thermal equilibrium density matrix at inverse temperature β0, and

A1µ and A2µ are external sources for the two copies of the currents on the SK contour, J1µ

and J2µ, respectively. The P in front of the exponential indicates path-ordering on the SK µ contour. Notice that the conservation of the currents ∂µJs = 0 for s = 1, 2 implies that the

4Previously, we used J to indicate sources that appeared in our generating functional. Now we are using Jµ to indicate quantum fields. 5Ordinarily, Noether’s theorem implies that such symmetries would lead to conserved currents, but in non-equilibrium systems, the relationship between symmetries and conserved quantitates is not so straight- forward. We will discuss how Noether’s theorem can fail in Chapter 3.

15 generating functional is invariant under the transformation

W[A1µ, A2µ] → W[A1µ + ∂µλ1, A2µ + ∂µλ2], (1.24)

for arbitrary functions λ1(x) and λ2(x) that vanish at spacetime infinities. We therefore see that W enjoys two gauge symmetries.

Letting ϕ1 and ϕ2 represent the IR degrees of freedom, we wish to construct an effec- tive action such that, when these IR fields are integrated out, we obtain the generating functional W. Schematically, we have

¹ W[A1µ,A2µ] iI [ϕ1,ϕ2;A1µ,A2µ] e = D[ϕ1ϕ2]e EFT . (1.25)

As of now, we know nothing of the form of the non-equilibrium effective action, IEFT. For- tunately, the gauge symmetries of W allow us to conveniently identify the IR degrees of

freedom. In particular, suppose that ϕ1 and ϕ2 are Stückleberg fields in the sense that they always appear in the forms

B1µ = A1µ + ∂µϕ1, B2µ = A2µ + ∂µϕ2. (1.26)

Then we have ¹ W[A1µ,A2µ] iI [B1µ,B2µ] e = D[ϕ1ϕ2]e EFT . (1.27)

Notice that a transformation of the form Asµ → Asµ + ∂µλs may, on the r.h.s., be absorbed into a redefinition of ϕs for s = 1, 2. As a result, when ϕs are integrated out, we are left with a generating functional W that is invariant under (1.24). It is all well and good that we have Stückleberg fields in the IR, but a discerning reader may wonder why there cannot be other fields in the IR. After all, there is nothing from the above discussion that excludes such a possibility. The answer is simple: we assume that the only IR degrees of freedom are those associated with either conserved Noether currents or

16 spontaneously broken symmetries. Since the U(1) charge is the only conserved quantity and 6 no symmetries are spontaneously broken, we are left with only ϕ1 and ϕ2 as the IR degrees of freedom. We define the “off-shell hydrodynamic” currents by

µ δIEFT µ δIEFT J1 ≡ , J2 ≡ − . (1.28) δA1µ δA2µ

We immediately see that the equations of motion for ϕ1 and ϕ2 are precisely the conservation µ µ equations for J1 and J2 , respectively. Thus far, our treatment has been quite general. It could describe systems at zero or finite temperature and could describe the U(1) charge of ordinary fluids or superfluids. To ensure that the system exists at finite temperature, we will need to impose the dynamical KMS symmetries and to ensure that the U(1) charge is diffusive (i.e. associated with an ordinary fluid as opposed to a superfluid) we will impose a set of time-independent gauge symmetries. Because our action describes the U(1) charge of an ordinary fluid, we must require that the charge not be spontaneously broken. It turns out that the correct way to describe an unbroken charge is by imposing the so-called “chemical shift” symmetries by

ϕ1 → ϕ1 + λ(xì), ϕ2 → ϕ2 + λ(xì), (1.29)

for arbitrary time-independent λ. Notice that because ϕ1 and ϕ2 are continuous on the

SK contour, we have, in the infinite future, ϕ1(+∞) = ϕ2(+∞). Therefore since λ is time independent, we must shift the two copies of ϕ by the same λ. Transforming to the retarded- advanced basis, we have

ϕr → ϕr + λ(xì), ϕa → ϕa, (1.30)

6Note: Lorentz boosts are explicitly broken.

17 where 1 ϕr = (ϕ1 + ϕ2), ϕa = ϕ1 − ϕ2. (1.31) 2

We will see later in this chapter that the retarded fields behave as classical variables and the advanced fields account for thermal and quantum fluctuations. By imposing the chemical shift symmetries, we ensure that the effective action can only depend on the retarded fields

in the form µ = ∂t ϕr , which can be interpreted as the chemical potential, which is consistent with the goal of describing the U(1) charge of an ordinary fluid. Notice that if we did

not impose the chemical shift symmetry then ϕr would be the Goldstone associated with a spontaneously broken U(1) charge.

Next, to ensure that our action exists at inverse temperature β0, we impose the dynamical KMS symmetries. For simplicity, we will work in the classical limit. Let Θ be a time-reversing symmetry. We will impose parity and charge inversion invariance so all choices of Θ will be equivalent. Under the classical dynamical KMS symmetries, we have

Br µ(x) → ΘBr µ(x), Baµ = ΘBaµ + iβ0Θ∂t Br µ(x). (1.32)

Assuming homogeneity (translational invariance) and isotropy (rotational invariance), we arrive at the leading-order Lagrangian

σ 2 LEFT = i Bai + χBa0Br0 − σBai∂t Bri, (1.33) β0 where σ ≥ 0 7 and χ are constants that must be fixed by experiment. Setting the sources to zero, Ar µ = Aaµ = 0, we have the equations of motion

χµÛ = σ∇2 µ, (1.34)

7This condition is mandated by the fact that the imaginary parts of non-equilibrium effective actions must always be non-negative.

18 where µ = ∂t ϕr is the chemical potential. We therefore see that the chemical potential satisfies the diffusion equation with diffusion constant D ≡ σ/χ. The requirement that equilibrium be stable forces χ ≥ 0.

1.6 Relativistic hydrodynamics

1.6.1 The standard approach

In the standard approach to relativistic hydrodynamics, we begin by identifying the conserved Noether currents. Because we are considering a relativistic system, we must impose Poincaré symmetry. As a result, we have the conserved, symmetric stress-energy tensor T µν. Next, it is often the case that particle number is conserved, so we suppose that there exists a conserved U(1) Noether current J µ. Thus the hydrodynamic equations are simply µν µ ∂µT = 0, ∂µJ = 0. (1.35)

These equations, however, are not sufficient to specify the theory. In particular, in d + 1 dimensions, we have only d + 2 conservation equations, while T µν has (d + 2)(d + 1)/2 independent components and J has d + 1 independent components. Evidently, more input is needed. To describe a fluid, we assume that locally, thermodynamic equilibrium is approximately realized. In this way, we can represent the stress-energy tensor and U(1) current in terms of local thermodynamic parameters and their derivatives. In equilibrium, there is a preferred rest frame, namely the zero-momentum frame. We therefore suppose that our Noether currents can depend on the unit-vector field uµ(x). We interpret uµ(x) as the local four- velocity of the fluid. Working in the rest-frame, the equilibrium density matrix is given by − 1 (H−µN) e T − 1 (H−µN) ρ = , Z = Tr[e T ], (1.36) Z

19 where H is the Hamiltonian and N is the Noether charge given by

¹ ¹ H = dd x Tˆ00, N = dd x Jˆ0. (1.37)

We see therefore, that once the rest-frame is known, the equilibrium state is fixed by T and µ, which are the equilibrium temperature and chemical potential, respectively. Supposing that we splash the liquid, we expect that there should be some approximate notion of lo- cal temperature and chemical potential. We therefore promote these variables to generic functions of the spacetime coordinates, T(x) and µ(x). The aim is now to formulate constitutive relations, which are equations that relate the Noether currents to the local equilibrium variables and their derivatives. The relationship is quite general and can depend on arbitrary numbers of derivatives. If, however, we are only interested in the IR behavior of fluids, then we may organize our constitutive relations in terms of a derivative expansion. That is, the more derivatives a term in the constitutive relations has, the less important is. At leading order in the derivative expansion, we have

µν µ ν µν T0 = (T, µ)u u + p(T, µ)∆ , (1.38) µ µ J0 = n(T, µ)u , where ∆µν = ηµν + uµuν and , p, and n are generic functions of T and µ and have the interpretation as local energy density, pressure, and particle number density, respectively. We use the subscript 0 on the Noether currents to indicate that they are the zeroth order terms in the derivative expansion. Notice that these constitutive relations are dictated by symmetry: they are the most generic expressions possible at leading order in the derivative expansion. At first order in the derivative expansion, the meanings of our local thermodynamic fields becomes ambiguous. In particular, the four-velocity, temperature, and chemical potential of a fluid are only properly defined in equilibrium. For example one may freely define a local

20 temperature field T0(x) that differs from the original T(x) by gradients of hydrodynamic variables. Consider the set of infinitesimal transformations

T → T0 = T + δT,

µ → µ0 = µ + δµ, (1.39)

uµ → u0µ = uµ + δuµ.

µ µ 0µ 0 Notice that δu uµ = 0 because u uµ = u uµ = −1. We can see that there is enough freedom in the definition of uµ that by redefining the fluid four-velocity, we may ensure that no charge flows in the fluid frame defined by uµ. This choice of four-velocity is known as the Eckhart frame. Alternatively, we can choose the four-velocity such that no energy flows in the rest frame of the fluid, This choice is known as the Landau frame. In this section, we will work in the Landau frame. Notice that we still have not fixed the definitions of T and µ. To do so we demand that there be no higher-order corrections of  and n as we continue to higher orders in the derivative expansion. There may, however, be higher-order corrections to p. The first-order corrections to the conserved currents in Landau frame are then

µν µν µν T1 = −ησ − ζ∆ θ, (1.40) µ µν µν J1 = −σT∆ ∂ν(µ/T) + χT ∆ ∂νT,

where σµν is the shear tensor and θ is the expansion scalar given by

  µν µα νβ 2θ µ σ = ∆ ∆ ∂αuβ + ∂βuα − ηαβ , θ = ∂µu . (1.41) d + 1

Imposing constraints from the second law of thermodynamics, we find that η, ζ, and σ are all non-negative and χT = 0.

21 1.6.2 The fluid worldvolume

At finite temperature the effective action is often not defined on the physical spacetime. Instead, we must introduce the notion of the so-called fluid worldvolume. To see why this is so, we will reproduce the arguments presented in [2]. Consider a fluid with no conserved currents other than the stress-energy tensor. The sources for the stress-energy tensor are the

metric tensors gsµν, where s = 1, 2 indicates on which leg of the SK contour the metrics live. Then, the SK generating functional takes the form

W[g1µν,g2µν]  †  e = tr U(+∞, −∞; g1µν)ρU (+∞, −∞; g2µν) , (1.42)

where U(+∞, −∞; gsµν) is the time-evolution operator in the presence of source gsµν. Since the

stress-energy tensor is conserved, the generating functional W[g1µν, g2µν] must be invariant µ under two independent diffeomorphism transformations. Let ζs (x) for s = 1, 2 represent two different diffeomorphisms. Then we have

ζ1 ζ2 W[g1µν, g2µν] = W[g1MN, g2MN ], (1.43)

ζs where gsMN for M, N = 0, 1, 2, 3 denote diffeomorphism transformations of gsµν. More ex- µ ν ζs ( ) ≡ ∂ζs ( ( )) ∂ζs plicitly, we have gsMN σ ∂σM gsµν ζs σ ∂σN . Now, we can use the Stückelberg trick and promote the gauge transformations to dynamical fields. In particular, we ‘integrate in’ the µ fields Xs (σ) such that the generating functional becomes

¹ W[g1µν,g2µν] iI [G1MN,G2MN ] e = D[X1X2] e EFT , (1.44) where µ ν ∂Xs ∂Xs GsMN ≡ gsµν(Xs(σ)) (1.45) ∂σM ∂σN

22 Physical spacetime (2) Fluid worldvolume Physical spacetime (1) µ µ X2 (σ) σM X1 (σ)

Figure 1.1: This figure depicts how the fluid world-volume with coordinates σM is mapped µ µ into two copies of the physical spacetime by the maps X1 (σ) and X2 (σ). The red and blue lines in the right-and left-hand panels are the images of the red and blue lines in the middle µ µ panel under the maps X1 (σ) and X2 (σ). are pull-back metrics. Notice that when performing the Stückelberg trick, we had to intro- duce the coordinates σM for M = 0, 1, 2, 3. We will refer to σM as ‘fluid coordinates’ and the corresponding manifold on which they live as the ‘fluid worldvolume.’ Then the fields µ Xs (σ) are the dynamical fields and describe the embedding of the fluid worldvolume into the physical spacetime. In a certain sense, it is always possible to ‘integrate in’ these Stückelberg

fields as long as W[g1µν, g2µν] is diffeomorphism invariant. However, there is no guarantee that the resulting EFT will be non-trivial; for example, if the equilibrium density matrix is µ a pure, zero-temperature ground state without any SSB then the fields Xs are pure gauge and hence have no dynamics. It turns out that to ensure the system exists in fluid phase, we must require that the fluid coordinates enjoy partial diffeomorphism symmetries:

• Fluid elements are indistinguishable from one another, so we should be able to freely relabel them in a time-independent manner. We therefore require invariance under

σI → σ0I(σJ), (1.46)

for I, J = 1, 2, 3.

23 • Treating each volume element as a point particle, we require independent world-line reparametrization invariance,

0 ∂ f 0 σ0 → σ0 = f (σ0, σI), τ(σ0, σI) → τ(σ0 , σI), (1.47) ∂σ0

where τ(σ0, σI) plays the role of the worldline einbein for the fluid volume element at σI. We can then gauge-fix τ = 1, in which case we have the residual symmetry

0 σ0 → σ0 (σI), (1.48)

for I = 1, 2, 3.

The partial diffeomorphism symmetries (1.46) and (1.48) can be succinctly encapsulated by the transformation law σM → σM + ξ M (σI), (1.49)

where ξ M (σI) are arbitrary functions of the spatial fluid coordinates σI for I = 1, 2, 3. The justifications provided above for imposing these diffeomorphism symmetries are merely in- cluded to build the reader’s intuition; it is not understood from first principles how they arise, only that they are necessary to describe systems in fluid phase. See Appendix A for an alternative perspective on the origins of these symmetries.

1.6.3 The action approach

If we wish to construct the effective action describing a relativistic fluid with conserved particle number, then we must formulate our effective action on the fluid worldvolume with the appropriate diffeomorphism gauge symmetries. To ensure conservation of energy, we µ include the fields Xs (σ) and to ensure conservation of charge, we include the fields ϕs(σ) that we encountered previously when constructing the diffusion action. Moreover, these fields must be Stückleberg fields and therefore our action may only depend on them in the

24 following packages

µ ν µ ∂Xs ∂Xs ∂Xs (σ) GsMN ≡ gsµν(Xs(σ)) , BsM (σ) = Asµ(Xs(σ)) + ∂M ϕs. (1.50) ∂σM ∂σN ∂σM

Then, to ensure that the U(1) particle-conserving symmetry is unbroken, we impose the time-independent chemical shift symmetries

ϕ1 → ϕ1 + λ(σì ), ϕ2 → ϕ2 + λ(σì ). (1.51)

Notice that now these symmetries are time-independent in the sense that they have no dependence on the time coordinate σ0 of the fluid worldvolume. Following [2], we will take the classical limit. To do so, it is helpful to restore factors of ~. In order to have a finite limit as ~ → 0, we must consider advanced fields to implicitly carry a factor of ~ relative to advanced fields, so we have

~ ~ ~ ~ g1µν = gµν + gaµν, g2µν = gµν − gaµν, A1µ = Aµ + Aaµ, A2µ = Aµ − Aaµ, (1.52) 2 2 2 2 and

µ µ ~ µ µ µ ~ µ ~ ~ X = X + X , X = X − X , ϕ1 = ϕ + ϕa, ϕ1 = ϕ − ϕa. (1.53) 1 2 a 2 2 a 2 2

Taking the ~ → 0 limit, we find that the physical spacetime diffeomorphism symmetries are

α β µ µ ∂X ∂X X → f (X), gµν → gαβ (1.54) ∂ f µ ∂ f ν

and µ µ µ Xa → Xa + fa (X), gaµν → gaµν − L fa gµν. (1.55)

where f and fa are arbitrary functions of X and Lw is the Lie derivative along a vector w.

25 µ Notice that only X coordinates enjoy diffeomorphism symmetry; Xa enjoy a residual transformation that is not a full diffeomorphism symmetry. As a result, we may perform a coordinate transformation and define the action on the physical spacetime coordinates x µ ≡ X µ. Then the fluid coordinates become dynamical fields σM (x). We are now in a position to identify the covariant building-blocks. The advanced building- blocks are

Gaµν(x) ≡ gaµν + LXa gµν, Baµ(x) ≡ Aaµ + ∂µϕa. (1.56)

M M To construct the retarded building-blocks, define Kµ (x) = ∂µσ (x) and let

µ µ −1 µ β ≡ βu ≡ (K )0, (1.57)

µ µ where u uµ ≡ −1. We identify β as the local inverse temperature field and u as the local fluid four-velocity field. Additionally, we have

µ µ = u ∂µϕ, (1.58) which we interpret as the local chemical potential field. At leading-order in the derivative expansion, these are the full set of covariant building-blocks. In the classical limit, the dynamical KMS symmetry transformation’s action on the re- tarded fields are equivalent to the action of Θ, the anti-unitary, time-reversing operator. On the advanced fields, we have

ρ gaµν → Θ[gaµν − iβ ∂ρgµν], (1.59) ρ Baµ → Θ[Baµ − iβ ∂ρBµ].

Notice that the advanced field transformations involve derivatives of the retarded fields. We therefore consider advanced fields as containing one more derivative than their retarded counterparts. As a result, the terms of the leading-order Lagrangian must be linear in

26 advanced fields. The most general form for such an Lagrangian is

1 µν µ L1 = T Gaµν + J Baµ, (1.60) 2 0

with µν µ µ µν µ µ T0 = (β, µ)u u + p(β, µ)∆ , J = n(β, µ)u , (1.61) where ∆µν = gµν + uµuν. Imposing the dynamical KMS symmetries gives the relations

∂p ∂p ∂p  + p = −β + µ , n = . (1.62) ∂β ∂µ ∂µ

Notice that these equations are equivalent to the first law of thermodynamics if we interpret  as the energy density, p as the pressure, and n as the number density. The next-to-leading- order Lagrangian takes the form

1 µν µ i µναβ µν i µνα L2 = T Gaµν + J Baµ + W GaµνGaαβ + iZ BaµBaν + Y Gaµν Baα. (1.63) 2 1 1 4 0 0 2 0

The most general expression for this action is quite complicated. Fortunately, we can em- ploy the generalized Landau-frame field-redefinitions of [3] to greatly simplify matters. In particular, it is possible to defined the fields such that

µν µ µναβ µν µνα uµT1 = uµJ1 = uµW0 = uµZ0 = Y0 = 0. (1.64)

We therefore have

µν µν ρ µν µ −1 µν T1 = −ησ − ζ∇ρu ∆ , J1 = −σβ ∆ ∇ν(βµ), (1.65)

27 where σµν is given by (1.41) and

  µναβ 1 µν W = 2β−1η ∆α(µ∆ν)β − ∆µν∆αβ + β−1ζ∆µα∆νβ, Z = β−1σ∆µν. (1.66) 0 d 0

We interpret η and ζ as the bulk and shear viscosities, respectively. And

 + p2 κ = σβ (1.67) n is the heat conductivity. All of these parameters are generic functions of β and µ.

We have constructed the terms of L2 to be invariant under the KMS symmetries (up to total derivatives). Interestingly, the coordinate transformations necessary to construct generalized Landau frame are not compatible with the dynamical KMS symmetries, so there is no reason to expect higher-order terms in this Landau-frame Lagrangian to respect the dynamical KMS symmetries. The resultant generating functional would, however, obey the standard KMS symmetries. Further, by happy coincidence, at leading order in dissipative quantities, the dynamical KMS symmetries hold exactly in generalized Landau frame [2].

µν µν µν Finally, notice that the stress-energy tensor T = T0 + T1 and the particle number µ µ µ current J = J0 + J1 constructed above take the same form as those constructed in §1.6.1. In this way, we have reproduced the standard theory for relativistic hydrodynamics using effective field theory principles.

1.7 The meaning of retarded and advanced fields

In this section, we will investigate the physical meanings of the r-and a-type fields. In particular, we will see that r-type fields play the role of classical field configurations, whereas a-type fields encode information about noise due to thermal and quantum fluctuations.

Suppose that we have a non-equilibrium effective action IEFT[ϕr, ϕa]. Notice that the third unitarity constraint of (1.16) requires that IEFT vanish if we set ϕa = 0, meaning that every term in the effective action must contain at least one a-type field. It turns out that the terms

28 in the effective action that are linear in a-type fields play very different roles from those with higher powers of a-type fields. Therefore, let us write

¹ 4 IEFT[ϕr, ϕa] = d x E[ϕr ]ϕa + I2[ϕr, ϕa], (1.68)

where I2 contains terms that are at least quadratic ϕa. Next, define η[ϕr, ζ] by

¹  ¹  −η[ϕr,ζ] 4 e ≡ D[ϕa] exp iI2[ϕr, ϕa] + i d xζϕa . (1.69)

Using this definition, we have

¹ ¹  ¹  iIEFT[ϕr,ϕa] 4 D[ϕr ϕa] e = D[ϕr ϕaζ] exp i d x (E[ϕr ] − ζ)ϕa − η[ϕr, ζ] ¹ (1.70) −η[ϕr,ζ] = D[ϕr ζ] δD[E[ϕr ] − ζ]e ,

where δD is the Dirac delta functional. For any given ζ, the argument of the delta functional has a straightforward interpretation; it contains the equations of motion, namely

E[ϕr ] = ζ. (1.71)

But we see from the r.h.s. of (1.70) that for each ζ, the delta functional is weighted by e−η[ϕr,ζ]. We therefore interpret ζ as a stochastic field with probability distribution e−η[ϕr,ζ].

1.8 Organization of the thesis

In this thesis, we will investigate properties of many-body systems out of finite tem- perature equilibrium. We will employ novel mathematical techniques, many of which are borrowed from high-energy physics, to aid in our analysis. The outline of the remaining chapters is as follows: In Chapter 2, we will investigate leading-order effective actions for finite-temperature

29 fluids, superfluids, solids, supersolids, nematic liquid crystals, and smectic liquid crystals in phases A, B, and C. With the exception of the fluid and superfluid actions, all other effective field theories are entirely new and are constructed with the aid of a novel coset construction that has been engineered to formulate non-equilibrium effective actions. This new coset construction postulates Goldstone-type modes associated with every symmetry, including unbroken symmetries. The distinguishing feature among broken and unbroken symmetries is the emergence of time-independent gauge redundancies associated with unbroken symme- tries. In addition to the usual inverse Higgs constraint, we identify two new such constrains that arise when unbroken Goldstones are present. In Chapter 3, we investigate higher-form symmetries in non-equilibrium systems and we explain in greater detail how symmetries can fail to have associated Noether currents. We find an additional inverse Higgs constraint that allows for the removal of second sound in cer- tain situations. We extend the non-equilibrium coset construction to account for higher-form symmetries and use it to construct a number of effective actions for various physical sys- tems. We exploit an already known relationship between ’t Hooft anomalies and spontaneous symmetry breaking to formulate new effective actions for magnetohydrodynamics in which a Legendre transform-type duality exists between the already-existing magnetohydrodynamics action and the new action we construct.

30 Chapter 2: The coset construction for non-equilibrium effective

actions

When trying to understand a many-body system, the goal is often not to keep track of all of the degrees of freedom; there are simply too many. Instead, the aim is to focus only on quantities that persist over long distances and extended time scales. To do this, one coarse-grains over the microscopic, or UV degrees of freedom to obtain an effective theory of the macroscopic, or IR observables. Often, one of the most important guiding principles in constructing such effective theories are symmetries and their corresponding Noether currents. For example the standard formulation of hydrodynamics comes from the equations for conservation of energy, momentum, and particle number [43]. Even though such an approach is heavily based in symmetry considerations it is not very systematic as other non-symmetry constraints must be imposed. In particular, it relies on the somewhat ill- defined notion of ‘near-equilibrium’ to construct constitutive relations among the conserved quantities and local thermodynamic parameters.1 Once this is done, the second law of thermodynamics must be imposed by hand, putting constraints on various coefficients. While this method has led to very powerful and general statements about hydrodynamic systems, there remain some difficulties. In particular, it is not at all clear what are the rules of the game for constructing hydrodynamic theories. For example, one might wonder if there are additional constraints beyond the second law of thermodynamics that one must impose. Additionally, a way to extend hydrodynamic theories to methodically account for thermal fluctuations or quantum effects is not at all clear. An approach that utilizes symmetries as the only input could resolve some of these issues. 1Near-equilibrium thermodynamics in the usual formulation is ill-defined in the sense that thermodynamic parameters like temperature and chemical potential are only rigorously defined in true equilibrium.

31 In order to make the approach to hydrodynamics—and many-body physics as a whole— more systematic, over the last decade or so there has been a great effort to reformulate hydrodynamics from the point of view of an effective action [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28]. The reasons for doing so are as follows. Effective actions provide powerful tools for describing systems in high-energy and particle physics because they can be constructed from symmetry principles alone. Once the symmetries and field content are specified, one constructs the effective action by writing down a linear combination of all terms that are consistent with the symmetries. Thus, there is no guesswork when formulating effective actions. One can then express local thermodynamic quantities like temperature and chemical potential in terms of the fields, yielding a more rigorous understanding of the concept of near-equilibrium thermodynamics. Applying these methods to condensed matter systems has provided powerful tools of analysis. From the EFT perspective, many-body systems can often be understood as systems that spontaneously break spacetime symmetries. In fact, it is often possible to classify the state of matter of a many-body system by its SSB pattern alone [1]. Then, the long-distance and late-time dynamics are described entirely by Goldstone bosons. For our purposes, this is very good news: there is a powerful and systematic way to construct EFTs of Goldstones known as the coset construction [44, 45, 46, 47, 48]. This construction takes the symmetry-breaking pattern—which is specified by the state of matter of the system—as the only input. The coset construction has been used to formulate numerous EFTs describing various states of matter [11]. Historically, the EFT approach to many-body physics has been unable to account for statistical fluctuations and dissipation. The reason is that ordinary actions describe noiseless, conservative systems and are therefore incapable of accounting for stochastic and dissipative dynamics. However, this changed with recent work [2, 3, 4, 5, 6, 7, 8] and [18, 19, 20, 21, 22, 23] in which a new kind of non-equilibrium action was formulated using the in-in formalism on the SK contour. Using symmetry principles alone these new non-equilibrium

32 EFTs can account for both statistical fluctuations as well as dissipation. Because these EFTs are constructed on the SK contour, the field content is doubled. In addition to the global symmetry group, there exist emergent gauge symmetries at finite temperature, the origins of which remain poorly understood [2]. Nevertheless, despite significant advances, in the current literature there exist very few systems for which the non-equilibrium EFTs are known. The reason is that no overarch- ing understanding of the field content, emergent gauge symmetries, or symmetry-breaking pattern has yet been provided for non-equilibrium systems. In this paper, we address these problems by combining two powerful tools: the non-equilibrium EFT approach to many- body systems and the coset construction.2 We term the result the non-equilibrium coset construction. Our approach is as follows: We start from the observation that at finite temperature, there exist Goldstone-like excitations corresponding to each symmetry generator even if it is not spontaneously broken. However, the Goldstones corresponding to the unbroken generators behave very differently than those arising from SSB. In particular the Goldstones corresponding to unbroken generators have infinitely many gauge symmetries analogous to the chemical shift symmetries of [9]. These gauge symmetries lead to diffusive behavior. Since the non-equilibrium effective action is defined on the SK contour, the field content is doubled. This requires the introduction of two cosets; one for each leg of the SK contour. We use these cosets to construct building-blocks for the non-equilibrium effective actions that transform covariantly under both the global symmetries as well as the chemical shift-type gauge symmetries. To demonstrate the validity of this new coset construction, we use it to formulate actions for already known EFTs, namely those of fluids and superfluids at finite temperature. Then, to demonstrate the utility of our approach, we construct novel EFTs for solids, supersolids, and four phases of liquid crystals, all at finite temperature. We thus vastly expand the 2The coset construction was recently used in [24] to build simple EFTs on the SK contour; however their approach only dealt with internal symmetries.

33 number of known non-equilibrium EFTs. Moreover, our formalism can be used to construct EFTs for essentially any state of matter that is describable by Goldstone (or Goldstone-like) excitations.

2.1 The zero-temperature coset construction: a review

Consider a relativistic quantum field theory whose full symmetry group is G, which includes both internal symmetries as well as the Poincaré group. Suppose that the ground state of the system spontaneously breaks the symmetry group G to the subgroup H. Then the IR dynamics of the system are described by Goldstone modes. If only internal symmetries are spontaneously broken, then for every broken symmetry generator in the coset G/H, there is a corresponding Goldstone boson. However, if spacetime symmetries are spontaneously broken, there are often fewer Goldstones than broken symmetry generators [44, 45, 46]. Often, the only gapless modes in the system are the Goldstones. If we are only concerned with the deep IR dynamics, we can integrate out all non-Goldstone modes to obtain an effective action for the Goldstones. It turns out that there is a systematic method for writing down Goldstone effective actions known as the coset construction. This section will give a brief review of the coset construction for spontaneously broken internal and spacetime symmetries. For in-depth discussions of the coset construction for spontaneously broken internal symmetries, consult [47] and for spontaneously broken spacetime symmetries, consult [48]. The coset construction is an especially powerful technique since it provides a systematic method for generating effective actions of Goldstone bosons using the SSB pattern as the only input. Supposing the symmetry-breaking patten is G → H, we represent the symmetry

34 generators by

P¯µ = unbroken translations,

TA = other unbroken generators, (2.1)

τα = broken generators,

where the generators τα and TA may be some combination of internal and spacetime gen- erators and we have assumed that there exist some notions of spacetime translations that remain unbroken. In this way, states can still be classified according to the corresponding notions of energy and momentum [1]. Importantly, we do not require that the unbroken

generators P¯µ be the original Poincaré translation generators (represented by Pµ); instead they can be some linear combination of Pµ and internal symmetry generators [1]. It will turn out that although P¯µ and TA both refer to unbroken generators, they will play very different roles in the coset construction. Therefore, it is convenient to define the subgroup H0 ⊂ H

that is generated exclusively by TA. The EFT of the Goldstones must be invariant under the full symmetry group G. While the action of the unbroken symmetry subgroup H is linearly realized on the fields, the action of the broken coset G/H is nonlinearly realized. In general, therefore, generic symmetry transformations will act in a highly non-trivial manner on the Goldstones. As a result writing down the most general effective action consistent with symmetries can be rather challenging unless we use the construction that follows. We are interested in ensuring that the effective action remain invariant under all symme- try transformations. The symmetries that act linearly are easy to deal with, but those that act non-linearly present more of a challenge. By definition, spontaneously broken generators act non-linearly on the fields while unbroken generators act linearly. However, unbroken translations act non-linearly on the coordinates, meaning that the coset of symmetry gen- 3 erators that have some sort of non-linear action is G/H0. It is convenient to parameterize

µ ¯ 3This is a somewhat hand-wavy justification for including eix Pµ in the coset. But in any case it turns

35 this coset by µ ¯ α γ(x, π] = eix Pµ eiπ (x)τα . (2.2)

Then, up to normalization, πα(x) correspond to the Goldstones.

By the definition of the coset G/H0, it is possible to express any element of G as the product of an element of the coset and an element of H0. As a result, for any element g ∈ G, we may write 0 0 g · γ(x, π] = γ(x , π ]· h0(x, π, g], (2.3)

0 0 where h0 is some element of H0 and x and π are transformed coordinates and fields. It is important to note that for given π and g, the terms on the r.h.s. of (2.3) can be explicitly computed if we know the commutation relations among the generators. We can therefore read off the transformations of the coordinates and Goldstones under an arbitrary symmetry transformation from (2.3). In particular, under the transformation by g, we have x → x0 and π → π0. Using the parameterization (2.2), we can construct the Maurer-Cartan one-form g−1dg. This one-form has the property that it can always be expressed as a linear combination of the symmetry generators [47, 48], so we have4

−1 ν  ¯ α A  g ∂µg = iEµ Pν + ∇νπ τα + Bν TA . (2.4)

Using the symmetry algebra alone, it is possible to compute the coefficients of each generator

α in the above expression. It can be checked that ∇µπ transforms covariantly under the full symmetry group G; we therefore will refer to it as the covariant derivative of π. Additionally,

A Bν transforms like a connection (i.e. gauge field) and we can therefore use it to compute out that (2.2) has the correct symmetry properties for our purposes. 4Here, E, ∇π and B are simply names for the terms that appear in the Maurer-Cartan form. We will see momentarily why we have chosen these particular labels.

36 higher-order covariant derivatives by

H −1 ν A ∇µ = (E )µ∂ν + iBµTA. (2.5)

ν And finally, Eµ serves as a vierbein; in particular, the invariant integration measure is d4x det E. It should be noted that the indices µ, ν need not be contracted in the usual way if the Lorentz group is broken. −1 After computing g ∂µg, it is then easy to identify the covariant building-blocks. At

α leading order in the derivative expansion we have ∇µπ ; and higher-order-derivative terms H α H H α are given by ∇µ (∇νπ ), ∇µ ∇ν ∇ρπ , etc. Then, the symmetry-invariant terms of the effective action are simply constructed by taking manifestly H-invariant combinations of these covariant terms. If boosts are broken but rotations are not, then examples of such

α β α β invariant terms are ∇0π ∇0π and ∇iπ ∇iπ , where repeated indices are summed over.

2.1.1 Inverse Higgs

We mentioned earlier that when only internal symmetries are broken, the number of Goldstones exactly matches the number of broken generators. However, when spacetime symmetries are broken, this need not be the case. In this section, we will see how this works from the perspective of the coset construction. Pragmatically, the rules of the game are as follows: Suppose that the commutator between an unbroken translation generator P¯ and a broken generator τ0 contains another broken generator τ, that is [P¯, τ0] ⊃ τ. Suppose further that τ and τ0 do not belong to the same

irreducible multiplet under H0. Then it turns out that it is consistent with symmetry transformations to set the covariant derivative of the τ-Goldstone in the direction of P¯ to zero. This gives a constraint that relates the τ0-Goldstone to derivatives of the τ-Goldstone, allowing the removal of the τ0-Goldstone. The setting of this covariant derivative to zero is known as an inverse Higgs constraint.

37 The physical reasons for imposing these inverse Higgs constraints have been investigated in [10, 46, 49]. Essentially, there are two possibilities. Firstly, sometimes when spacetime symmetries are spontaneously broken, the resulting Goldstones do not correspond to inde- pendent fluctuations. As a result, there can be multiple Goldstone field configurations that all correspond to the same physical state. From this viewpoint, the inverse Higgs constraints can be understood as a convenient choice of gauge-fixing condition. Secondly, when a Gold- stone can be removed via inverse Higgs constraints, it is often the case that if we did include the Goldstone in the effective action, it would have an energy gap. But we are often only interested in gapless excitations, so we may integrate out these fields. From this perspective, the inverse Higgs constraints correspond to integrating out gapped Goldstones. For the pur- poses of this paper, with the exception of §2.3.1, we will take a purely pragmatic approach and always impose inverse Higgs constraints whenever possible, remaining agnostic about the precise reasons for doing so.

2.1.2 Zero-temperature superfluids: a simple example

Thus far, our discussion of the coset construction has been rather abstract. This section will focus on a simple concrete example: the coset construction of the zero-temperature superfluid EFT. First, since our theory is relativistic, it ought to be Poincaré-invariant. In our ‘mostly plus’ convention, the Poincaré algebra is

i[Jµν, Jρσ] = ηνρJµσ − ηµρJνσ − ησµJρν + ησν Jρµ,

i[Pµ, Jρσ] = ηµρPσ − ηµσ Pρ, (2.6)

i[Pµ, Pν] = 0,

where Pµ are the translation generators and Jµν are the Lorentz generators. From the EFT perspective, a superfluid is defined as a system that has a conserved U(1) charge Q such that both Q and P0 (i.e. time translation) are spontaneously broken but a diagonal subgroup,

P¯0 ≡ P0 + µQ is preserved [50, 51]. Physically, Q is the charge associated with particle-

38 number conservation and µ is the chemical potential. Additionally, since every condensed matter system (including superfluids) has a zero-momentum frame, boosts are necessarily

broken. As a result, the broken generators are the U(1) charge Q and Lorentz boots Ki ≡ J0i.

The unbroken translations are P¯0 = P0 + µQ and P¯i = Pi for i = 1, 2, 3 and the remaining

1 i jk unbroken generators are the spatial rotation generators Ji ≡ 2  Jjk. With this symmetry- breaking pattern, the coset is then

µ ¯ i g = eix Pµ eiπ(x)Qeiη (x)Ki . (2.7)

By explicit computation, the Maurer-Cartan form is

−1 ν  ¯ i i  g ∂µg = iEµ Pν + ∇νπQ + ∇νη Ki + Ων Ji , (2.8) where

ν ν Eµ = Λµ ,

−1 ν 0 ∇µπ = (E )µ∂νψ − µδµ, (2.9) i −1 ν −1 0i ∇µη = (E )µ[Λ ∂νΛ] , 1 Ωi = (E−1)ν ijk[Λ−1∂ Λ]jk, µ 2 µ ν

i µ iη Ki µ i jk k where Λ ν ≡ [e ] ν and ψ ≡ µt + π. Here,  Ωµ can be thought of as the spin-connection (or at least the spatial components of it) and transforms as a gauge field under rotations. Now, notice that the commutator between unbroken spatial translations and broken

boosts yields [P¯i, Kj] ⊃ iδi j µQ, meaning that we may impose inverse Higgs constraints to remove the boost Goldstones. In particular, we may set to zero the covariant derivative of π along the i = 1, 2, 3 directions, yielding

µ 0 = ∇iπ = Λ i∂µψ. (2.10)

39 These constraints can be solved to give a relation between the boost Goldstones and deriva- tives of the U(1) Goldstones i η ∂iψ tanh η = − . (2.11) η ∂0ψ

Using these relations, we can remove the boost Goldstones as dynamical degrees of freedom.

Thus, at leading order in the derivative expansion, the only covariant building-block is ∇0π.

By plugging (2.11) into the expression for ∇0π, we find that

q µ ∇0π = −∂µψ∂ ψ − µ. (2.12)

Since µ is just a constant, we find that at leading order in derivatives, the zero-temperature EFT must be built out of an arbitrary function of

q µ y ≡ −∂µψ∂ ψ. (2.13)

Since det E = 1, we have that the invariant integration measure is just d4x, so the leading- order action is ¹ S = d4x P(y), (2.14) where P is some function that is determined by the equation of state. Terms that are higher

i i order in the derivative expansion can be constructed using ∇µη as well as Ωµ, which plays the role of a connection and can be used to create covariant derivatives of the form (2.5). But to keep things as streamlined as possible, in this chapter we will only construct leading-order actions.

2.2 The non-equilibrium coset construction

Before we are able to state the procedure for constructing non-equilibrium effective ac- tions with the method of cosets, we must first understand the IR field content of condensed matter systems at finite temperature. It turns out that Goldstone and Goldstone-like excita-

40 tions at finite temperature are a bit different than at zero temperature. After the field content is established, we will outline the procedure for using cosets to construct non-equilibrium effective actions, which we term the non-equilibrium coset construction.

2.2.1 Goldstones at finite temperature

At finite temperature, SSB occurs when a symmetry generator fails to commute with the equilibrium density matrix. Goldstone’s theorem tells us that for every spontaneously broken symmetry, there is a corresponding Goldstone mode (unless inverse Higgs constraints are imposed). However, at finite temperature, there are other excitations that survive over long distances and extended time scales, which resemble Goldstones. Our claim is that there are in fact Goldstone-like fields for every symmetry regardless of whether it is broken or unbroken (unless inverse Higgs-type constraints are imposed). This claim is best understood from a semi-classical perspective. Semi-classically, the density matrix is a purely pragmatic tool that is used to account for our classical ignorance of the true micro-state of the system. Essentially every (classical) micro-state in a thermal statistical ensemble corresponds to a highly chaotic classical field configuration, which will in general spontaneously break every symmetry. As a result, we expect that the non-equilibrium effective action should consist of Goldstones as if every symmetry of the theory were spontaneously broken. We will refer to the Goldstones corresponding to spontaneously broken generators as broken Goldstones and those corresponding to unbroken generators as unbroken Goldstones. It turns out that there is an important distinction between broken and unbroken Gold- stones. In particular, in every known case, unbroken Goldstones possess infinitely many gauge symmetries. For specific examples, consult [2, 3, 4, 9, 52]. While it is not fully un- derstood from first principles where these gauge symmetries come from, Appendix A gives a pragmatic explanation for why they should exist. We will take as a well-motivated assump-

tion that these gauge symmetries act as follows. Let TA be the unbroken generators and let

A s (σ) be the unbroken Goldstones. Here, s = 1, 2 indicates on which leg of the CPT the

41 fields live. Then we have the gauge redundancies

A( ) A( ) A( I ) eis σ TA → eis σ TA eiλ σ TA, (2.15)

for arbitrary spatial functions λA(σI). We use the convention that M, N = 0, 1, 2, 3 and I, J = 1, 2, 3. Additionally, the effective action will be invariant under a certain subgroup of diffeomorphisms that act on the fluid worldvolume coordinates σM by [2]

σM → σM + ξ M (σI), (2.16)

which we interpret as additional gauge symmetries. Since it is unknown how exactly these symmetries arise, it is conceivable that they may not hold in all situations. However, we will see in the following sections that (2.15) and (2.16) are in fact the correct symmetries for a wide range of physical systems.

2.2.2 The method of cosets

As discussed in the previous subsection, at finite temperature there is a Goldstone mode corresponding to each symmetry of the theory (unless inverse Higgs-type constraints are imposed). As a result, in the coset construction we must parameterize the full symmetry group of the theory with Goldstone modes.5 Since unbroken Goldstones enjoy a gauge symmetry whereas broken Goldstones do not, we must distinguish between the two types of Goldstones. Additionally, we assume that there exist some sorts of unbroken translation 5Since we are parameterizing the full symmetry group as opposed to merely the non-linearly realized coset, it is not technically correct to call our construction a coset construction. Nevertheless we will use the term ‘coset construction’ for the sake of linguistic continuity.

42 generators [1]. The global symmetry generators of the theory are as follows:6

P¯µ = unbroken translations,

TA = other unbroken generators, (2.17)

τα = broken generators.

Finally, let G be the full symmetry group of the theory (including spacetime symmetries), let H ⊂ G be the unbroken subgroup, and let H0 ⊂ H be the subgroup generated by TA. We will construct our theory on the fluid worldvolume with coordinates σM for M = 0, 1, 2, 3. Parameterize an arbitrary group element by

µ α A iX (σ)P¯µ iπ (σ)τα i (σ)TA gs(σ) = e s e s e s , (2.18)

where s = 1, 2 indicates on which leg of the SK contour the fields live. Each gs for s = 1, 2 transforms under the same global symmetry action. Notice that the spacetime coordinates x µ that appear in the ordinary coset construction (2.2) have been promoted to dynamical µ fields Xs (σ) in the non-equilibrium coset construction. These fields serve to embed the fluid worldvolume into the physical spacetime. We are interested in building-blocks for the non-equilibrium effective action that can be constructed from the Maurer-Cartan one-form and that transform covariantly under the global symmetry group G as well as under the gauge transformations

M M I gs(σ) → gs(σ + ξ (σ )), (2.19) A I iλ (σ )TA gs(σ) → gs(σ)e ,

6When using the coset construction for theories with gauge symmetries, there are two main approaches. In [11, 53], gauge transformations are treated as if they are physical symmetries and generators for each gauge transformation appear in the coset. In [48], gauge symmetries are treated as redundancies and only global symmetry generators appear in the coset. Both methods yield correct results, but we will use the second approach as it is far simpler.

43 for arbitrary spatial functions ξ M (σI) and λ(σI).7 To this end, we will treat the coefficients of unbroken generators of H0 in the Maurer-Cartan form as gauge-fields that have the partial gauge-invariance defined by the second line of (2.19). The Maurer-Cartan one-form is

−1 µ ¯ α  A gs ∂M gs = iEsM Pµ + ∇µπs τα + iBsMTA, (2.20)

µ α where EsM are the vierbeins, ∇µπs are the covariant derivatives of the broken Goldstones, A 8 and certain components of BsM behave like gauge connections. These objects transform under (2.19) as follows. Under the TA-gauge transformations, the vierbeins, the covariant

A A derivatives, and Bs0 transform linearly, whereas BsI transform as

A A −1 −1 BsITA → BsI h0 · TA · h0 − ih0 · ∂I h0, (2.21)

A I I iλ (σ )TA A where h0(σ ) ≡ e ; that is, BsI transform as connections. Under the fluid diffeo- µ α A morphism transformations, Es0, ∇µπs , and Bs0 all transform as scalars, whereas the spatial component of the vierbeins and the connections transform as

∂σM E µ → E µ , sI ∂σ0I sM (2.22) ∂σM B A → B A , sI ∂σ0I sM

where σ0M ≡ σM + ξ M (σI). The building-blocks that transform covariantly under both the global symmetries and the gauge symmetries (2.19) are as follows: First, there are the building-blocks from the

α usual coset construction, namely ∇µπs , which transform covariantly under (2.19), and to

7Once again, we use M, N = 0, 1, 2, 3 and I, J = 1, 2, 3. 8 −1 Notice that terms from log(g1 · g2) are invariant under the global symmetry group G and covariant under the gauge transformations, yet they are not used as building-blocks of the effective action. Our claim is that only building-blocks that arise from the Maurer-Cartan one-form are permissible. We will see explicitly that this is the case in the next chapter.

44 take higher-order covariant derivatives we can use

∂ H A , ∇ = ∂I + iB TA, (2.23) ∂σ0 I rI

A 1 A A  where BrI = 2 B1I + B2I and I = 1, 2, 3. To contract coordinate indices, we use the metrics

µ ν GsMN = EsM ηµν EsN . (2.24)

Second, there are new building-blocks that involve the unbroken Goldstone degrees of free-

µ A dom, namely Es0 and Bs0, which transform covariantly. Finally, we have terms that involve µ −1 M A A A combinations of 1-and 2-fields. Notice that E1M (E2 )ν and BaM ≡ B1M − B2M transform µ ν covariantly and we can contract coordinate indices with E1M ηµν E2N . After computing these building-blocks, inverse Higgs-type constraints can be imposed to remove extraneous Goldstone modes. The basic idea behind inverse Higgs constraints is that we may set to zero any objects that transform covariantly under the symmetries and gauge symmetries as long as the resulting equations can be algebraically solved for one set of Goldstones in favor of derivatives of another set of Goldstones. We will see in subsequent examples that there are three kinds of inverse Higgs-type constraints. In the next chapter, we will find a fourth inverse Higgs-type constraint that is responsible for removing second sound. There are the usual inverse Higgs constraints that exist for zero-temperature systems. Suppose that the commutators between an unbroken translation generator P¯ and a broken generator τ0 contains another unbroken generator τ, that is [P¯, τ0] ⊃ τ. Further, suppose that 0 τ and τ do not belong to the same irreducible multiplet under H0. Then, it is consistent with symmetry transformations to set the covariant derivative of the τ-Goldstone in the direction of P¯ to zero. This gives a constraint that relates the τ0-Goldstone to derivatives of the τ-Goldstone, allowing the removal of the τ0-Goldstone. Notice that the ordinary inverse Higgs constraints only allow for the removal of certain

45 broken Goldstones by relating them to derivatives of other broken Goldstones. One might wonder if there are other possibilities. Perhaps we could remove broken Goldstones by relating them to derivatives of unbroken Goldstones; or perhaps we could specify some constraints that allow us to remove unbroken Goldstones. It turns out that both of these are possible; the rules are as follows:9

• Thermal inverse Higgs: Suppose that at finite temperature, the commutator between

a broken generator τ and the unbroken time-translation generator P¯0 contains an un-

broken spacetime translation generator P¯, that is [τ, P¯0] ⊃ P¯. Then we may set to µ ¯ zero the component of E0 in the direction of P. This gives an equation that can be algebraically solved to yield an expression for the τ-Goldstone in terms of derivatives of the P¯-Goldstone. This equation allows the removal of the τ-Goldstone.

• Unbroken inverse Higgs: Suppose that at finite temperature, the commutator between an unbroken generator T and an unbroken spacetime translation generator P¯0 con- tains another unbroken spacetime translation generator P¯, that is [T, P¯0] ⊃ P¯. Con-

µ µ −1 M sider the matrix A ν ≡ (E1)M (E2 )ν , where M = 0, 1, 2, 3 are coordinate indices and

µ, ν = 0, 1, 2, 3 are Lorentz indices. Then we may set to zero the components of Aµν in the directions of P¯ and P¯0. Suppose that under the dynamical KMS symmetry trans-

formation, Aµν → A˜µν. Then, we may also set to zero the components of A˜µν in the directions of P¯ and P¯0. These conditions give constraints that relate the T-Goldstone to derivatives of the P¯-Goldstone, allowing the removal of the T-Goldstone.10

One might wonder if the aforementioned inverse Higgs constraints are the only possible such constraints. From a very general stand-point, an inverse Higgs constraint can be im- posed any time there is a covariant object that when set to zero, results in an equation which 9To get a better understanding of why these new inverse Higgs constraints can be imposed, see Ap- pendix C. In this appendix, to keep things as concrete as possible, we perform an in-depth analysis of inverse Higgs-type constraints for fluids without conserved charge. 10It turns out that the unbroken inverse Higgs constraints do not necessarily remove the unbroken Gold- stones from the covariant building-blocks. However they do remove the unbroken Goldstones entirely from the invariant building-blocks and hence from the effective action; see Appendices C.2-C.3.

46 can be algebraically solved for one set of Goldstones in terms of another. With this level of generality, it is hard to say if we have been exhaustive.11 However, most inverse Higgs con- straints are related to commutator relations of the form [A, P¯] ⊃ B, for unbroken translation generator P¯ and some generators A and B. Because the set of unbroken generators forms a closed algebra, it is impossible for B to be broken yet A unbroken. Every other combination of broken and unbroken, however, is permitted. If both A and B are broken, then we have ordinary inverse Higgs constraints; if A is broken and B unbroken, we have thermal inverse Higgs constraints; and if both A and B are unbroken, we have unbroken inverse Higgs con- straints. Thus, our list of inverse Higgs constraints is exhaustive if we assume that all such constraints must be related to commutator relations.12 Once inverse Higgs constraints have been imposed and an effective action has been con- structed, it is then necessary to impose the dynamical KMS symmetry. Often, at leading order in the derivative expansion (which is the only order to which we will be working in the subsequent examples in this chapter) the only effect of the dynamical KMS symmetry will be to mandate that the effective action factorize as

3 IEFT[χr, χa] = SEFT[χ1] − SEFT[χ2] + O(χa ), (2.25)

1 3 where χr ≡ 2 (χ1 + χ2), χa ≡ χ1 − χ2, and O(χa ) counts as higher order in the derivative expansion. The only exceptions to this rule that appear in this paper are the EFTs for nematic and smectic C phases of liquid crystals. When this rule holds, it allows us to work

α with just one copy of the fields. Thus, at leading order, the building-blocks ∇µπ describe

µ A the broken Goldstones and E0 and B0 describe the unbroken Goldstones. For the sake of simplicity, in the subsequent sections, we will construct EFTs for various states of matter at leading order in the derivative expansion. As a result, we will construct ordinary actions 11In fact we will see that there is an additional inverse Higgs constraint that can be imposed in certain circumstances. But it takes a bit more mathematical machinery to make sense of this fourth constraint. 12Strictly speaking, the thermal and unbroken inverse-Higgs constraints make the assumption that B is an unbroken translation generator. It is possible to conceive of more general possibilities for B, but such inverse Higgs constraints will not be relevant for any constructions in this paper.

47 with just one copy of the fields whenever possible. It should be noted that the primary reason for defining effective actions on the SK contour is to account for dissipation. When an action factorizes according to (2.25), the leading-order physics is fully captured by an ordinary action and is thus non-dissipative. Therefore, the majority of the effective actions constructed in this paper will not account for dissipative effects. However, our coset construction gives a clear, almost mechanical prescription for constructing higher-order terms, which if included in the EFT, will account for dissipation. We leave this important work of computing higher-order terms with the coset construction for future study.

2.3 Fluids and superfluids at finite temperature

To demonstrate that our non-equilibrium coset construction presented in the previous section gives correct results, we will reproduce the known effective actions for fluids and su- perfluids at finite temperature. Along the way, we will find that finite-temperature framids— systems for which only boosts are spontaneously broken [1]—can be thought of as fluids. In particular, at finite temperature, the boost Goldstones automatically become gapped and can therefore be integrated out, resulting in the ordinary fluid EFT. Finally, the equations of motion for fluids and superfluids from the EFT perspective have been studied in great detail in [2, 9, 12]. Thus, we will not study them here. For all other states of matter, however, we will include brief discussions of the resulting equations of motion.

2.3.1 Fluids

Consider a fluid without a conserved charge; the symmetry group is just the Poincaré group, whose algebra is given by (2.6). The unbroken generators are Pµ for translations and

1 i jk Ji ≡ 2  Jjk for spatial rotations; the broken generators are Ki ≡ J0i for boosts. Therefore,

48 the most general group element is

µ i i g(σ) = eiX (σ)Pµ eiη (σ)Ki eiθ (σ)Ji . (2.26)

The resulting Maurer-Cartan one-form is

−1 µ  i  i g ∂M g = iEM Pµ + ∇µη Ki + iΩM Ji, (2.27) where the vierbein, covariant derivative, and spin-connection are given by

µ ν µ EM = ∂M X [ΛR]ν ,

i −1 M −1 0j ji ∇µη = (E )µ [Λ ∂M Λ] R , (2.28) 1 Ωi = i jk[R−1Λ−1∂ (ΛR)]jk, M 2 M

i i i j iθ (σ)Ji i j µ iη (σ)Ki µ such that R = [e ] and Λ ν = [e ] ν. Because spatial rotations are unbroken, the action must be invariant under the transformations

g(σ) → g(σM + ξ M (σI)), (2.29) i I g(σ) → g(σ)· eiλ (σ )Ji,

M I i I for arbitrary spatial functions ξ (σ ) and λ (σ ). Notice that [Pi, Kj] = iδi j P0, meaning that we may impose the thermal inverse Higgs constraints, allowing the removal of ηi. In

i particular E0 transforms covariantly under both the global symmetry group G as well as the gauge and diffeomorphism symmetries (2.19). Therefore we may fix

µ i ∂X j ji 0 = E = Λµ R . (2.30) 0 ∂σ0

49 i j µ i Since R is invertible, this is equivalent to ∂0X Λµ = 0, from which we find that

ηi ∂ Xi − 0 tanh η = t , (2.31) η ∂0X

p where η ≡ ηiηi. For a more detailed calculation, see Appendix C.1. Now we can impose unbroken inverse Higgs constraints to remove the rotation Gold- stones.13 In order to do this, we must recall that the field-content is doubled. Notice that

[Pi, Jj] = iijk Pk, so we may set

−1 M −1 M (E1)iM (E2 )j − (E1)jM (E2 )i = 0, (2.32)

which gives

√ ∂X µ −1 T −1 ij −1 iν 1 j R1 · R2 = M · M · M , M ≡ (Λ2 ) ν (Λ1)µ . (2.33) ∂X2

See Appendix C.2 for more detailed calculations. To get an intuitive sense of what is hap- pening, let us expand the above equation to linear order in the fields. Transforming from the 1,2-basis to the r, a-basis (1.18), we have

ì ì ì θa = ∇ × Xa, (2.34)

where ∇×ì represents the ordinary curl and not a covariant derivative. Then, performing the (classical) dynamical KMS transformation on both sides gives

∂θìr ∂Xìr = ∇ì × . (2.35) ∂σ0 ∂σ0

But notice that the linearized version of (2.29) implies that our action must be invariant

13At this level in the derivative expansion, it is not actually necessary to solve the unbroken inverse Higgs constraints, but we will do so just to demonstrate that it can be done. When imposing the unbroken inverse Higgs constraints in later sections, such calculations will be omitted.

50 ì ì ì I ì under θr (σ) → θr (σ) + λ(σ ), meaning that at linear order, θr can only appear in the ì 0 ì effective action in the form ∂θr /∂σ . As a result, (2.35) is sufficient to remove θr as an ì ì independent degree of freedom. Thus, θr and θa have been successfully removed from the ì effective action, so no rotational Goldstones remain. To see how θr can be removed at the non-linearized level, see Appendix C.3. Thus the only covariant building-block at leading order in the derivative expansion is

t q µ ν 14 15 E0 = −E0 ηµν E0 . Defining the metric by

µ ν µ ν ∂X ∂X GMN ≡ E ηµν E = ηµν , (2.36) M N ∂σM ∂σN

t 2 we find that the leading-order action is a generic function of G00 = −(E0) and is given by

¹ √ 4 SEFT = d σ −GP(G00). (2.37)

Notice that the above system has the same symmetry-breaking pattern as a framid ex- cept for the fact that it exists at finite temperature [1]. Thus, we can interpret a finite- temperature framid as a fluid. With this interpretation, the meaning of the thermal inverse

i Higgs constraint is as follows. If we did not set E0 = 0, then we could use it as a covari- i i i ant building-block. But E0 ⊃ η + ··· , meaning that terms involving η with no derivatives must exist in the action. Therefore ηi has an energy gap.16 Thus at finite temperature, framid Goldstones necessarily develop a gap and can therefore be integrated out. It should be noted, however, that at sufficiently low temperatures, the framid Goldstones’ energy gap may be quite small. In this case it would only be appropriate to integrate them out if we

14 0 To avoid ambiguity, instead of writing E0 , we replace the Lorentz index with a t. This way it is clear t that E0 has one raised Lorentz index and one lowered coordinate index. Whenever we feel that there may be an ambiguity, we will use µ = t instead of µ = 0 to indicate time-like components of Lorentz vectors and tensors. 15The metric tensor defined in (2.36) agrees exactly with the pull-pack fluid metrics of [2]. 16At finite temperature, the notion of an energy gap is ill-defined. Really, what we mean is that the propagator will die off in space and time exponentially fast. But since the leading-order hydrodynamic action takes the form of an ordinary action with an energy gap, we use the term ‘energy gap.’

51 are interested exclusively in the very deep IR behavior of the system. Now suppose that the fluid carries a conserved U(1) charge Q and exists at finite chemical 17 potential. Then the unbroken translations are P¯0 = P0 + µQ and P¯i = Pi. Only boosts are spontaneously broken and the most general group element is

µ ¯ i i g(σ) = eiX (σ)Pµ eiπ(σ)Qeiη (σ)Ki eiθ (σ)Ji . (2.38)

The resulting Maurer-Cartan one-form is

−1 µ  ¯ i  i g ∂M g = iEM Pµ + ∇µη Ki + iBMQ + iΩM Ji, (2.39) where the vierbein, covariant derivative, and spin-connection are given by (2.28) and the U(1) gauge field is given by t BM = ∂M ψ − µEM, (2.40) where ψ(σ) ≡ µXt(σ) + π(σ). We are interested in constructing an action that is invariant under (2.29) as well as the chemical shift gauge symmetry

I g(σ) → g(σ)· eλQ(σ )Q, (2.41)

I i for arbitrary spatial function λQ(σ ). As before, we may remove η by imposing thermal inverse Higgs constraints (2.30), and we may remove θi by imposing unbroken inverse Higgs constraints (2.32).

t Notice that the U(1) gauge field B0 transforms as a scalar. Since E0 also transforms as t a scalar, it is convenient to define the invariant building-block B0 ≡ B0 + µE0 = ∂0ψ. The

17 Notice that both P0 and Q are unbroken, so defining P¯0 = P0 + µQ reflects a choice since any linear combination of P0 and Q is unbroken. But it is the most natural choice given that this definition of P¯0 allows ¯0 ¯0 us to express the equilibrium density matrix in the usual form ρ = e−β0 P /tr e−β0 P .

52 effective action at lowest order in derivatives is therefore

¹ √ 4 SEFT = d σ −GP(B0, G00). (2.42)

Notice that (2.37) and (2.42) are precisely the leading-order actions presented in [2, 3, 4]. They appear to differ from the actions presented in [9], but it can be shown that they are equivalent; see Appendix D.

2.3.2 Superfluids

Consider a superfluid at finite temperature. In addition to possessing Poincaré symme-

try (2.6), such a system must also have a conserved U(1) charge Q such that P¯0 = P0 + µQ, ¯ 1 Pi = Pi, and Ji = 2 ijk Jjk are the unbroken generators and Q and Ki = J0i are the broken generators [1, 12]. The most general group element is

µ ¯ i i g(σ) = eiX (σ)Pµ eiπ(σ)Qeiη (σ)Ki eiθ (σ)Ji . (2.43)

And the resulting Maurer-Cartan one-form is

−1 µ  ¯ i  i g ∂M g = iEM Pµ + ∇µπQ + ∇µη Ki + iΩM Ji, (2.44) where

µ ν µ EM = ∂M X [ΛR]ν ,

−1 M 0 ∇µπ = (E )µ ∂M ψ − µδµ, (2.45) i −1 M −1 0j ji ∇µη = (E )µ [Λ ∂M Λ] R , 1 Ωi = i jk[R−1Λ−1∂ (ΛR)]jk, M 2 M

i 0 µ iη (σ)Ki µ such that ψ(σ) ≡ µX (σ) + π(σ), the Lorentz boost matrix is Λ ν(σ) = [e ] ν, and k the rotation matrix is Rij(σ) = [eiθ (σ)Jk ]ij. To remove the boost Goldstones, impose inverse

53 ρ −1 M µ µ Higgs constraints 0 = ∇iπ = Λ i(e )ρ ∂M ψ, where eM ≡ ∂M X . These constraints give a relation between the boost Goldstones and the U(1) Goldstones

i −1 M η (e ) ∂M ψ tanh η = − i . (2.46) η −1 M (e )0 ∂M ψ

p MN Using this relation, we have that ∇t π = −G ∂M ψ∂N ψ − µ, giving our first symmetry- invariant building-block

p MN Y ≡ −G ∂M ψ∂N ψ. (2.47)

t 1 −1 M µ Equation (2.46) tells us that [ΛR] µ = Y (e )µ ∂M ψ, from which we find that E0 ∇µπ = µ [1 − Y ]∂0ψ, yielding our second symmetry-invariant building-block

∂ψ B0 ≡ . (2.48) ∂σ0

And just as in the fluid case, we have

µ ν G00 = E0 ηµν E0 (2.49) as our third symmetry-invariant building-block. Finally, following the same calculation as in the previous section, we can impose unbroken inverse Higgs constraints to remove the rotation Goldstones. Thus, the effective action describing a superfluid at finite temperature is, to lowest order in derivatives,

¹ √ 4 SEFT = d σ −GP(Y, B0, G00). (2.50)

This effective action looks somewhat different from the one presented in [12], but can be shown to be equivalent; see Appendix D. Further, in the zero-temperature limit, the effective action loses its dependence on X µ(σ), meaning that it only depends on the physical-spacetime version of Y, namely (2.13), reproducing the standard superfluid EFT (2.14).

54 2.4 Solids and supersolids at finite temperature

Now that we have established that the non-equilibrium coset construction can reproduce known results, we turn our attention to the construction of novel EFTs. The simplest physical systems for which the non-equilibrium effective actions are as of yet unknown are solids and supersolids. In this section, we will construct the leading-order non-equilibrium EFTs for these states of matter.

2.4.1 Solids

Consider a chargeless crystalline solid at finite temperature. In addition to possessing Poincaré symmetry (2.6), such a system must also have internal translation symmetry gener-

ators Qi for i = 1, 2, 3 that commute with each other. We then take P¯0 = P0 and P¯i = Pi + Qi 1 to be the unbroken generators such that Qi, Ki = J0i, and Ji = 2 ijk Jjk are the broken generators [1]. The most general group element is

µ ¯ i i i g(σ) = eiX (σ)Pµ eiπ (σ)Qi eiη (σ)Ki eiθ (σ)Ji . (2.51)

And the resulting Maurer-Cartan one-form is

−1 µ  ¯ i i i  g ∂M g = iEM Pµ + ∇µπ Qi + ∇µη Ki + ∇µθ Ji , (2.52) where

µ ν µ EM = ∂M X [ΛR]ν ,

i −1 M i i ∇µπ = (E )µ ∂M ψ − δµ, (2.53) i −1 M −1 0j ji ∇µη = (E )µ [Λ ∂M Λ] R ,

i −1 M ijk −1 −1 jk ∇µθ = (E )µ  [R Λ ∂M (ΛR)] ,

55 i i i i µ iη (σ)Ki µ such that ψ (σ) ≡ X (σ) + π (σ), the Lorentz boost matrix is Λ ν(σ) ≡ [e ] ν, and k the rotation matrix is Ri j(σ) ≡ [eiθ (σ)Jk ]i j. To remove the boost and rotation Goldstones, i ì 18 µ µ impose inverse Higgs constraints ∇t π = 0 and ∇ × πì = 0, respectively. Let eM ≡ ∂M X . Then by imposing the first inverse Higgs constraint, we find that

j η −1 M j −1 i j tanh η = −(e ) ∂M ψ (a ) (2.54) η t

i j −1 M j (i j) where a ≡ (e )i ∂M ψ . Imposing the second inverse Higgs constraint tells us that ∇ π = (Y 1/2)ij − δij, where i j MN i j Y ≡ G ∂M ψ ∂N ψ , (2.55)

such that GMN is the inverse of the metric (2.36). We therefore have our first set of invariant building-blocks. Notice that because there is no notion of unbroken rotational symmetry, Yi j is truly symmetry-invariant and not merely covariant. We have the additional invariant

µ i i j −1 i j j objects E0 ∇µπ = (δ −(Y ) )∂0ψ , from which we extract our next set of invariant building- blocks ∂ψi Zi ≡ . (2.56) ∂σ0

Lastly, we have the usual time-like piece of the fluid metric (2.36), G00, as our final building- block. Thus the effective action describing a crystalline solid at finite temperature is, to lowest order in derivatives,

¹ √ 4 i j i SEFT = d σ −GP(Y , Z , G00). (2.57)

Often, solids possess an additional unbroken U(1) symmetry corresponding to particle- number conservation. Let Q be the corresponding generator. Then, the unbroken transla-

18 ì ì Here, ∇ should be thought of as the spatial components of the covariant derivative ∇µ. Thus ∇ × πì is not the curl of πì.

56 tions are P¯0 = P0 + µQ and P¯i = Pi + Qi. The most general group element is

µ ¯ i i i g(σ) = eiX (σ)Pµ eiπ (σ)Qi eiη (σ)Ki eiθ (σ)Ji eiπ(σ)Q. (2.58)

The corresponding Maurer-Cartan one-form is just (2.52) with the addition of a term in-

t 0 volving the U(1) gauge-field BM = ∂M ψ − µEM for ψ(σ) = X (σ) + π(σ). As when we added

a conserved charge to fluids, we now have the additional building-block B0 ≡ ∂0ψ. Thus the effective action describing a crystalline solid with conserved U(1) charge at finite temperature is, to lowest order in derivatives,

¹ √ 4 ij i SEFT = d σ −GP(Y , Z , B0, G00). (2.59)

As a consistency check, let us take the zero-temperature limit. This amounts to removing all dependence on ∂X µ(σ)/∂σ0. We see therefore that the zero-temperature EFT only depends on the physical-spacetime version of Yi j, namely

i j i µ j y ≡ ∂µψ ∂ ψ , (2.60) which agrees with the results of [11]. To connect with other formulations of the finite-temperature hydrodynamics of solids [54, 55, 56, 57, 58], we compute the equations of motion. If we vary the chargeless action with

µ µν respect to X , we find the conservation of the stress-energy tensor ∂νT = 0. The stress- energy tensor is given by

T µν = εuµuµ + p∆µν + r µν, ∆µν ≡ ηµν + uµuν, (2.61) where ∂P 1 ∂X µ − − µ p = P, ε = 2G00 P, u = √ 0 (2.62) ∂G00 −G00 ∂σ

57 are the thermodynamic pressure, energy density, and four-velocity, respectively and

∂P i j rµν = ∂µψ ∂νψ (2.63) ∂Yij is the elastic stress-tensor. So far, our theory appears to agree with standard results. How- ever, we have the additional equations of motion that come from varying ψi,

iµ iµ µ j ∂P µ ∂P ∂µJ = 0, J ≡ 2∂ ψ + u . (2.64) ∂Yi j ∂Zi

These new equations indicate that the solid degrees of freedom can be excited without af- fecting the stress-energy tensor. The physical interpretation is that our solid exhibits second sound modes, analogous to those found in finite-temperature superfluids. The ordinary sound modes of solids consist of transverse and longitudinal vibrational modes of the lattice. However, if a finite-temperature crystalline solid has a sufficiently pristine lattice structure and Umklapp scattering events can be ignored, then there will be a bath of thermalized solid that behave as a through which an additional fluid-like longitudinal can propagate [59]. The additional equations of motion (2.64) account for the fact that the solid degrees of freedom can move independently of the phonon gas. Further, in the standard formulation, the thermoelastic variables depend on the temperature and a symmetric tensor,

−1/2 ij which we can identify with (−G00) /β0 and Y respectively, where β0 is the equilibrium inverse temperature. However, in our formalism, there is an additional quantity Zi. Just like the additional equations of motion, this quantity owes its existence to the presence of second sound. In particular, it is non-zero when the phonon gas flows with respect to the solid lattice. In the next chapter, we will see how second sound can be removed.

2.4.2 Supersolids

Consider an anisotropic supersolid at finite temperature. In addition to possessing Poincaré symmetry (2.6), such a system must also have internal translation symmetry gen-

58 erators Qµ for µ = 0, 1, 2, 3 that commute with each other such that P¯µ = Pµ + Qµ are the unbroken generators and Qµ and Jµν are the broken generators [1]. The most general group element is µ( ) ¯ µ( ) i µν( ) g(σ) = eiX σ Pµ eiπ σ Qµ e 2 θ σ Jµν . (2.65)

And the resulting Maurer-Cartan one-form is

  −1 µ ν 1 νλ g ∂M g = iE P¯µ + ∇µπ Qν + ∇µθ Jνλ , (2.66) M 2

where

µ ν µ EM = ∂M X Λν ,

ν −1 M ν ν ∇µπ = (E )µ ∂M ψ − δµ, (2.67)

−1 M −1 ∇µθνλ = (E )µ [Λ ∂M Λ]νλ,

µ µ µ µ iηi(σ)K such that ψ (σ) = X (σ)+π (σ) and the Lorentz transformation matrix is Λ ν(σ) = [e i ·

k iθ (σ)Jk µ e ] ν. To remove the Lorentz Goldstones, impose inverse Higgs constraints

ρ −1 M 0 = ∇[µπν] = Λ µ(e )ρ ∂M ψν − (µ ↔ ν). (2.68)

−1 M T Letting Mµν ≡ (e )µ ∂M ψν, the inverse Higgs constraints merely require that Λ · M be

a symmetric matrix, where factors of ηµν have been suppressed. Using this result, we have 1/2 that ∇(µπν) = (Y )µν − ηµν, where

µν MN µ ν Y = G ∂M ψ ∂N ψ (2.69)

are our first set of symmetry-invariant building-blocks and we use the convention that

1/2 ρλ 1/2 (Y )µρη (Y )λν = Yµν. (2.70)

59 ν µ µ −1 µ ν Additionally, we have E0 ∇νπ = (δν − (Y )ν )∂0ψ . Our next set of invariant building-blocks is therefore ∂ψ µ Z µ ≡ . (2.71) ∂σ0

And finally, we have the usual G00 building-block. Thus the effective action describing a crystalline supersolid at finite temperature is, to lowest order in derivatives,

¹ √ 4 µν µ SEFT = d σ −GP(Y , Z , G00). (2.72)

As in the solid EFT case, taking the zero-temperature limit amounts to removing all depen- dence on ∂X µ(σ)/∂σ0. Thus at zero temperature, the effective action can only depend on the physical-spacetime version of Y µν, namely

µν µ ρ ν y ≡ ∂ρψ ∂ ψ , (2.73) which agrees with the results of [11]. Finally, the equations of motion are very similar to those of solids. In particular if we replace the ψi for i = 1, 2, 3 with ψ µ for µ = 0, 1, 2, 3 in equations (2.61) and (2.64), we find

µν µν the finite-temperature supersolid equations of motion are ∂νT = 0 and ∂ν J = 0.

2.5 Liquid crystals at finite temperature

We now turn our attention to the construction of EFTs for more exotic states of matter, namely those of liquid crystals. There are myriad distinct phases of liquid crystals, so for the sake of brevity (and the reader’s patience) we will focus on four of the most common phases: nematic liquid crystals and smectic liquid crystals in phases A, B, and C. See Figure 2.1 for a graphical representation of various phases of liquid crystals. At low temperatures, the system exists in the crystalline solid phases, spontaneously breaking all spatial translational and rotational symmetries; the corresponding EFT is therefore given

60 Figure 2.1: This figure depicts the microscopic appearance of various phases of liquid crystal in order from lowest to highest temperature. The phases are (a) crystalline solid, (b) smectic liquid crystal in phase C, (c) smectic liquid crystal in phase A, (d) nematic liquid crystal, and (e) isotropic, i.e. fluid phase. Notice that (a) spontaneously breaks the most symmetries and each subsequent phase spontaneously breaks fewer and few symmetries until (e) only breaks boosts.

by (2.57) or (2.59). As the temperature rises, symmetries are restored until at high tem- peratures, the system exists in fluid phase in which no symmetries other than boosts are spontaneously broken; the corresponding EFT is therefore given by (2.37) or (2.42). The aim of this section will be to construct EFTs for the intermediate phases that partially break the ISO(3) group of spatial translations and rotations. Notice that in Figure 2.1, (a) spontaneously breaks all translations and rotations, corre- sponding to crystalline solid phase. Parenthetically, the profile of smectic liquid crystal in phase B appears almost identical to that of the solid; however in phase B, the horizontal lay- ers of molecules are allowed to slide past each other. Phases (b) and (c) both spontaneously break translations in the vertical direction, but not in the horizontal directions; therefore both exist in smectic liquid crystal phase. The difference between these two states of matter is that (b) spontaneously breaks all spatial rotations whereas (c) does not break rotations about the vertical direction; thus (b) depicts phase C and (c) depicts phase A. Phase (d) does not spontaneously break spatial translations in any direction, but does spontaneously break rotations about the horizontal axes, meaning that it exists in nematic liquid crystal

61 phase. Finally, (e) does not spontaneously break any spatial translations or rotations and is therefore in isotropic fluid phase.

2.5.1 Nematic liquid crystals

Liquid crystals in nematic phase are composed of oblong molecules that, like the molecules of an ordinary fluid, bounce around chaotically and therefore cannot form a lattice structure; however, on average the long axes of the molecules align, thereby breaking isotropy. Because no lattice structure can form, spacetime translations remain unbroken, but the aligned long axes of the molecules spontaneously break rotations. Non-relativistically, the order param- eter associated with broken rotations is a unit vector nì, pointing parallel to the long axes of the molecules. As a result the SO(3) symmetry group of spatial rotations is broken to SO(2) [60]. However, for any choice of nì, the long axes of the molecules could be equally 1 well specified by −nì. Therefore, it is common to use the order parameter Qi j ≡ ninj − 3 δij, which is invariant under the Z2 symmetry nì → −nì. To extend this order parameter to the

i j relativistic case, it is natural to define Qµν ≡ Qi j δµδν. We therefore see that boosts are spontaneously broken as well. Then, relativistically, the SSB pattern of the Poincaré group is ISO(1,3) → R2 × ISO(2). (2.74)

We also have the unbroken discrete Z2 symmetry associated with nì → −nì. Without loss of 19 generality choose hnìi = ˆz and let indices A, B = 1, 2. Then Pµ and J3 are the unbroken generators and JA and Ki are the broken generators. It will turn out that we need to use two copies of the fields to construct the leading-order action for this phase of matter. Parameterize the most general group elements by

µ i A 3 iX (σ)Pµ iη (σ)Ki iθ (σ)JA iθ (σ)J3 gs(σ) = e s e s e s e s . (2.75)

19The indices A, B are not to be confused with the indices indicating unbroken symmetry generators that we used in previous sections.

62 The resulting Maurer-Cartan one-forms are

−1 µ  A i  3 gs ∂M gs = iEsM Pµ + ∇µθs JA + ∇µηsKi + iΩsM J3, (2.76) such that

µ ν µ EsM = ∂M Xs [Λs Rs]ν ,

A −1 M AB3 −1 −1 B3 ∇µθs = (Es )µ  [Rs Λs ∂M (Λs Rs)] , (2.77) i −1 M −1 0j ji ∇µηs = (Es )µ [Λs ∂M Λs] Rs ,

3 −1 −1 12 ΩsM = [Rs Λs ∂M (Λs Rs)] ,

µ i ij A 3 iηs Ki µ iθs JA iθs J3 i j where Λs ν ≡ [e ] ν and Rs ≡ [e e ] . We can impose the thermal inverse Higgs constraints µ i ∂Xs j ji 0 = E = Λsµ R , (2.78) s0 ∂σ0 s

µ i ij which reduce to ∂0Xs Λsµ = 0 because Rs are invertible. Then, as in the ordinary fluid case, we have i i ηs ∂0Xs tanh ηs = − t , (2.79) ηs ∂0Xs √ p i i t where ηs ≡ ηsηs. Once again, Es0 = −Gs00. Additionally, we can impose the unbroken 3 inverse Higgs constraint, which removes the rotation Goldstone θs . Define the vector

k 1 ijk i −1 M A ≡  (E1) (E ) . (2.80) a 2 M 2 j

3 The unbroken inverse Higgs constraint is then Aa = 0. At this order in the derivative expansion it is not necessary to solve this inverse Higgs constraint, so in the interest of brevity we will not. In the case of ordinary fluids, we had the additional unbroken inverse Higgs constraints

A Aa = 0. But now JA are broken so we may not impose such constraints. We therefore have

63 A A a covariant building-block that mixes 1-and 2-fields, namely Aa . Next, define ∇θr implicitly as follows. Under a dynamical KMS transformation, we have20

A A A Aa → ΘAa + iΘβ0∇θr . (2.81)

The full set of covariant building-blocks at this order in derivatives is

A A A A Aa , ∇θr , ∇Bθs , ∇3θs , Gs00, (2.82)

µ ν where Gs00 ≡ Es0ηµν Es0. Thus, at leading order in the derivative expansion, the effective action is constructed from SO(2)×Z2-invariant combinations of the above terms. The con- tribution to the non-equilibrium effective action that does not contain mixtures of 1-and

2-fields is S1 − S2, where

¹ p 1 h [ B] S = d4σ −G 2P(G ) + K (G )(∇ Aθ )2 + K (G )(∇ θ A)2 s s 2 s00 1 s00 s 2 s00 A s (2.83) A 2i +K3(Gs00)(∇3θs ) .

A 2 Notice that we could have included a term involving (∇t θs ) , but we will see that the nematic A 2 A degrees of freedom are diffusive, so we will consider ∂0θs to count at the same order as ∂i θs in the derivative expansion. In the r, a-basis, the contribution to the non-equilibrium effective action that contains mixtures of 1-and 2-fields is

¹   4 p A A i A 2 Imix = d σ −Gr −M(Gr00)∇θr Aa + M(Gr00)(Aa ) , (2.84) β0

where β0 is the equilibrium inverse temperature and the relationship between the coefficients of the two terms is fixed by the (classical) dynamical KMS symmetry. Then, up to higher-

20 A A Notice that ∇θr involves a time-like derivative of θr .

64 order corrections, the full non-equilibrium effective action is given by

IEFT = S1 − S2 + Imix. (2.85)

To compare the above action built with the non-equilibrium coset construction to more

i i3 standard results, let the unit vector ns ≡ Rs indicate in which direction the elongated molecules of the nematic are oriented. Then, working in the classical, non-relativistic limit, we have

    A A i 1 j Û i i Û j j  i 1 k i i k k  ∇θ A = nÛ − n ∂j X − ∂ X n n − n ∂k X − ∂ X n (2.86) r a r 2 r r r r a 2 r a a r

and

[A B] 2 ì 2 A 2 ì 2 A 2 ì 2 (∇ θs ) = (∇ · nìs) , (∇Aθs ) = (nìs ·(∇ × nìs)) , (∇3θs ) = (nìs × (∇ × nìs)) , (2.87)

where the dot (Û) indicates differentiation with respect to σ0 and hence is the usual material

derivative when acting on nìr . Note that our EFT is invariant under nìs → −nìs. For simplicity, assume that the nematic and hydrodynamic modes decouple so we can focus just on the nematic degrees of freedom. Then, the equations of motion for the rotation Goldstones can be written in the form

¹ Û 1 δF 1 3  ì 2 ì 2 ì 2 nì = − , F ≡ d x K1(∇ · nì) + K2(nì ·(∇ × nì)) + K3(nì × (∇ × nì)) , (2.88) M δnì 2

where it is understood that M and Ki for i = 1, 2, 3 are evaluated on the equilibrium value of 1 G00. We therefore see that M F can be interpreted as the relaxation rate times the free energy, and K1, K2, and K3 as proportional to the standard splay, twist, and bend respectively [61]. We have thus recovered the standard nematic equations of motion.

A Notice that, linearizing in θr , the dispersion relation for the nematic degrees of freedom

65 is of the form ω ∝ ik2, indicating diffusion, which is an intrinsically dissipative effect. It is worth noting that this dissipation owes entirely to the fact that the action cannot be

factorized according to (2.25) because of the inclusion of the term Imix. Further, the necessity

of including Imix—and therefore the resulting diffusive dispersion relation—is an automatic consequence of the commutation relations of the Poincaré algebra; we did not need to assume diffusive behavior a priori. Finally, we often expect particle number to be conserved, requiring the inclusion of an additional unbroken U(1) charge Q. Let π be the corresponding unbroken Goldstone. Then just as in §2.3.1 and §2.4.1, we have the additional invariant building-block

∂ψ 0 B0 ≡ , ψ(σ) ≡ µX (σ) + π(σ). (2.89) ∂σ0

As a result, the effective action is identical to (2.85) except P, M, and Ki for i = 1, 2, 3 are

now functions of both G00 and B0.

2.5.2 Smectic liquid crystals

Smectic liquid crystals are composed of oblong molecules that form layers such that along one spatial direction, the liquid crystal has a periodic structure. Without loss of generality,

take P3 to be the spontaneously broken translation generator orthogonal to the layers. To ensure that some notion of unbroken translations along the ˆz direction exists, it is necessary

to introduce an internal translation symmetry generated by Q3 that is spontaneously broken

such that the diagonal subgroup generated by P¯3 ≡ P3 + Q3 is unbroken. Throughout this section, we will use A, B = 1, 2 to indicate spatial indices perpendicular to the ˆz direction.

Phase A

In phase A, the long axes of the molecules are, on the average, aligned with the ˆz direction. ¯ 3 Thus, Pµ = Pµ + δµQ3, and J3 are the unbroken generators and JA, Ki, and Q3 are the broken

66 generators. The most general group element is21

µ ¯ 3 3 A 3 A g(σ) = eiX (σ)Pµ eiπ (σ)Q3 eiη (σ)K3+iθ (σ)JA+iθ (σ)J3 eiη (σ)KA . (2.90)

The resulting Maurer-Cartan one-form is

−1 µ  ¯ 3 A i  3 g ∂M g = iEM Pµ + ∇µπ Q3 + ∇µθ JA + ∇µη Ki + iΩM J3, (2.91)

such that

µ ν µ EM = ∂M X [LΛ]ν ,

3 −1 M 3 3 ∇µπ = (E )µ ∂M ψ − δµ,

A −1 M AB3 −1 −1 B3 ∇µθ = (E )µ  [Λ L ∂M (LΛ)] , (2.92)

i −1 M −1 −1 0i ∇µη = (E )µ [Λ L ∂M (LΛ)] ,

3 −1 −1 12 ΩM = [Λ L ∂M (LΛ)] ,

A 3 A 3 3 3 3 µ iη KA µ µ iη K3+iθ JA+iθ J3 µ where ψ ≡ X + π and we have Λ ν ≡ [e ] ν and L ν ≡ [e ] ν. Begin by imposing unbroken inverse Higgs constraints; as in the previous subsection, they serve to remove the rotation Goldstone θ3. And once again, solving these unbroken inverse Higgs constraints is not necessary at this order in the derivative expansion. Next, impose ordinary

3 3 A 3 inverse Higgs constraints ∇t π = ∇Aπ = 0, which remove θ and η and tell us that

3 p MN 3 3 ∇3π = G ∂M ψ ∂N ψ − 1, (2.93)

yielding the first building-block

MN 3 3 Y ≡ G ∂M ψ ∂N ψ , (2.94)

21We use a somewhat non-standard parameterization of g so that the inverse Higgs constraints are easier to solve.

67 µ ν where the metric is given as usual by GMN = EM ηµν EN . Now impose thermal inverse Higgs constraints to remove ηA given by

µ A ∂X A 0 = E = [LΛ]µ . (2.95) 0 ∂σ0

Then, we have A µ A η ∂0X Lµ tanh η − , ⊥ = µ t (2.96) η⊥ ∂0X Lµ

p A A µ ν where η⊥ ≡ η η . This allows us to construct our final two building-blocks G00 = E0 ηµν E0 µ 3 and E0 ∇µπ , the second of which is just

∂ψ3 Z ≡ . (2.97) ∂σ0

Thus, the leading-order action for smectic liquid crystals in phase-A is

¹ √ 4 SEFT = d σ −GP(Y, Z, G00). (2.98)

To compare with known results, let us now expand this EFT to quadratic order in the fields. For simplicity, we will neglect the fluid degrees of freedom. Then, transforming to the physical spacetime, the quadratic Lagrangian describing the behavior of the smectic mode is

1 h i L(2) = − M (πÛ3)2 − M (∂ π3)2 − M (∂ π3)2 , (2.99) 2 0 1 A 3 3 for constants M0, M1, and M3. The resulting equations of motion,

3 M1 2 3 M3 2 3 πÜ = ∂Aπ + ∂3 π , (2.100) M0 M0 agree with the results of [62, 58]. For readers interested in the full equations of motion, they can be obtained by setting ψ1 = ψ2 = 0 in the solid equations of motion in §2.4.1.

68 Finally, if particle number is conserved, we include the additional building-block (2.89),

meaning that P in (2.98) may now depend on B0.

Phase B

Smectic liquid crystals in phase B, like in phase A, consist of vertically stacked layers that can slide past each other. In phase A, the molecules within a given layer are able to move around freely without forming a lattice structure. In phase B, however, the molecules in a given layer are locked in place. If we take the ˆz direction to be perpendicular to the stacked layers, then phase B is essentially a solid that cannot sustain uniform x-z and y-z shears [58]. As a result, the effective action is almost identical to (2.57)—or if it contains an unbroken U(1) charge it is almost identical to (2.59). The only difference is that to allow the layers to slide past each other, we must impose the symmetries

ψ A → ψ A + f A(ψ3), (2.101) where A = 1, 2 and f A is an arbitrary function of ψ3. Physically, these symmetries indicate that we can translate each layer in the x-y plane independently without changing the macro- scopic state of the system. At the level of the EFT, these additional symmetries mean that the effective action can only depend on Yij in the combinations

ij 11 33 13 2 22 33 23 2 33 b ≡ detY , b1 ≡ Y Y − (Y ) , b2 ≡ Y Y − (Y ) , Y , (2.102) and it can only depend on the i = 3 component of Zi. Finally, the equations of motion are just a special case of the solid equations of motion given in §2.4.1.

Phase C

Phase C is much like phase A except now the long axes of the molecules do not on average align with the ˆz direction, meaning that J3 is now spontaneously broken. Thus the

69 only difference between phases A and C is that phase C has a Goldstone associated with the 3 broken generator J3 denoted by θ . Further, this Goldstone cannot be removed with inverse Higgs-type constraints. Going through a similar procedure to the one in §2.5.2, we find that the effective action 3 has the same building-blocks as (2.98) with five invariant additions. First, we have ∇µθ , where µ = 0, 1, 2, 3 are now merely labels and do not need to be contracted in any particular 3 way since the entire Lorentz group is spontaneously broken. Additionally, we have Aa, which 3 is the ˆz-component of (2.80). The addition of the building-block Aa, which mixes 1-and 2-fields, means that the non-equilibrium effective action cannot factorize into the difference of ordinary actions. The contribution to the non-equilibrium effective action that does not contain mixtures of 1-and 2-fields is S1 − S2, where

¹   4 p 1 ij 3 3 Ss = d σ −Gs P(Ys, Zs, Gs00) + M (Ys, Zs, Gs00)∇iθ ∇j θ , (2.103) 2 s s

such that repeated indices i, j are summed over and Mi j is symmetric under exchange of i and j.22 Just as in the case of nematic phase, we will see that the rotation Goldstone, θ3 is 3 2 3 diffusive, so we will consider ∂0θ and ∂i θ to be the same order in the derivative expansion. 3 Define ∇θr implicitly as follows. Under a dynamical KMS transformation, we have

3 3 3 Aa → ΘAa + iΘβ0∇θr . (2.104)

In the r, a-basis, the contribution to the non-equilibrium effective action that contains mix- tures of 1-and 2-fields is

¹   4 p 3 3 i 3 2 Imix = d σ −Gr −M0(Gr00)∇θr Aa + M0(Gr00)(Aa) . (2.105) β0

22The sum over i and j is purely for notational convenience as i and j are merely labels.

70 As a result, up to higher-order corrections, the full non-equilibrium effective action is

IEFT = S1 − S2 + Imix. (2.106)

If particle number is conserved, we have the additional building-block (2.89). Then the non-equilibrium effective action is identical to (2.106), except the coefficient functions P,

Mi j, and M0 may now depend on B0. The full equations of motion for smectic C are quite complicated and therefore not terribly enlightening. Thus, for the sake of simplicity assume that the dynamics of θ3 decouple from those of X µ and π3. Then, the equations of motion for X µ and π3 are identical to those of the smectic A phase. And the linearized equation of motion for θ3 is, ignoring the complications of hydrodynamical fluctuations, ij Û3 M 3 θ = ∂i∂j θ . (2.107) M0

Thus, smectic C phase looks just like smectic A phase with the addition of a diffusive mode θ3, agreeing with standard results [58].

It is worth noting that the presence of the term Imix, as in the nematic phase, leads to a diffusive dispersion relation and hence dissipation. Thus nematic and smectic C phases of liquid crystal are the only two states of matter considered in this paper that exhibit dissipation at leading order in the derivative expansion.

2.6 Conclusion

In this chapter, we defined a systematic coset construction of non-equilibrium EFTs and used it to formulate both known and heretofore unknown non-equilibrium EFTs for condensed matter systems. We postulated that IR dynamics of thermal systems out of equilibrium can be characterized by Goldstones almost as if every symmetry of the system were spontaneously broken. However, the Goldstones corresponding to spontaneously bro- ken symmetry generators behave rather differently than those corresponding to unbroken

71 symmetry generators. In particular, the unbroken Goldstones possess infinitely many gauge symmetries, whereas the broken Goldstones, like ordinary Goldstones at zero temperature, have no such gauge symmetries. The approach of [11] was to treat these infinitely many symmetries as if they were true symmetries of the underlying theory. As a result, each of these symmetries required its own generator and Goldstone to parameterize the coset of broken symmetries.23 By contrast, our approach was to treat these additional symmetries as gauge redundancies in the style of [48], thereby circumventing the need to introduce infinitely many symmetry generators and Goldstones. Further, the coset construction presented in this chapter allows one to formulate non-equilibrium actions in full generality, whereas previous attempts at the coset construction for systems with spontaneously broken spacetime symmetries can only be used to formulate ordinary actions that are incapable of accounting for statistical fluctuations and dissipation.

23Since [11] only constructed actions to leading order in the derivative expansion, it was not actually necessary to introduce all of the infinitely many symmetry generators. However, extending to arbitrarily higher orders using their method would require the addition of infinitely many symmetry generators and Goldstones; see [53] for a general discussion of how gauge symmetries can be treated as genuine symmetries in the coset construction.

72 Chapter 3: Higher-form symmetries and ’t Hooft anomalies in

non-equilibrium systems

3.1 Introduction

In the previous chapter, we introduced an extension of the coset construction to formulate non-equilibrium actions. We were able to construct EFTs for a wide range of many-body systems using symmetry principles alone. While ordinary symmetries provide a powerful guiding principle for constructing EFTs, they are not the whole story. This is particularly true for non-equilibrium systems. In this chapter we focus on two main extensions of ordinary symmetry:

• There are many phases of matter that are characterized by topological features, known as topological phases of matter. It was long believed that these topological phases could not be classified according to symmetry principles alone; however, recent work has chal- lenged this idea [63]. In particular, it is suggested that a kind of generalized symmetry principle—known as higher-form symmetries—can be employed to describe topological features. In this way, topological properties of many-body systems can be classified according to Landau’s approach. The Noether charges corresponding to these higher- form symmetries describe the conservation of higher-dimensional extended objects. For example a p-form charge counts the number of charged p-dimensional objects. Such higher-form symmetries arise in many areas including gauge theories, magnetohydro- dynamics, dual theories of superfluids and solids, and many others [63, 64, 65, 66, 67, 68, 69, 70, 55, 52, 71, 72, 73, 74, 75]. In fact, a common feature of systems with spon- taneously broken U(1) symmetries—like superfluids and solids [29]—is the existence of an emergent higher-form symmetry with mixed ’t Hooft anomaly [76]. Further, when

73 a higher-form symmetry is spontaneously broken, a generalized Goldstone theorem guarantees the existence of a gapless mode.

• The role of symmetries is somewhat different for non-equilibrium systems than for systems at zero-temperature that are described by ordinary actions. In particular, for ordinary actions, Noether’s theorem furnishes a one-to-one correspondence between physical symmetries and conserved quantities. In non-equilibrium effective actions, however, no such one-to-one correspondence is guaranteed. There can be exact sym- metries of the action that have no corresponding conserved charge. We will see that these situations arise when there is a particular kind of obstruction to introducing gauge fields for a given symmetry. We therefore term this kind of non-conservation a non-equilibrium ’t Hooft anomaly.1

It is the aim of this chapter to systematically investigate how higher-form symmetries and non-equilibrium ’t Hooft anomalies can be used to construct non-equilibrium effective actions. To aid in the systematization of our approach, we employ the coset construction [29, 44, 45, 46, 47, 48, 35]. We find that for spontaneously broken p-form symmetries, the corre- sponding equivalent of the Maurer-Cartan form is no longer a one-form but is instead a p+1- form. When an ordinary or higher-form symmetry is spontaneously broken, a (generalized) Goldstone theorem guarantees the existence of a gapless mode [64]. For non-equilibrium systems, even unbroken symmetries can have corresponding gapless modes that resemble Goldstone excitations. At the level of the coset construction, these modes corresponding to unbroken symmetries can be accounted for by introducing certain gauge-redundancies [29]. We extend these gauge-redundancies to account for gapless modes corresponding to unbro- ken p-form symmetries. Further, we find that any action constructed exclusively using terms furnished by the Maurer-Cartan form will have a one-to-one correspondence between physi- 1We term such phenomena non-equilibrium ’t Hooft anomalies because they exists when a symmetry cannot be gauged; however, there are important differences with standard ’t Hooft anomalies. Usually, a ’t Hooft anomaly does not out-right kill the conservation of a given Noether current. Instead, any attempt to gauge a symmetry leads to non-conservation. Non-equilibrium ’t Hooft anomalies, by contrast, kill the conservation of the current even when no gauge fields have been introduced.

74 cal symmetries and conserved quantities. In order to destroy this one-to-one correspondence, we find that there is a Maurer-Cartan-like object that can be used to furnish terms that lead to non-equilibrium anomalies of the effective action. To build the readers’s intuition for higher-form symmetries and ’t Hooft anomalies, we investigate the non-equilibrium properties of finite-temperature electromagnetism in vari- ous settings at the level of quadratic effective actions. We use these various examples to illustrate key features of non-equilibrium systems: (1) whether or not a symmetry appears spontaneously broken depends on the time-scale over which the system is observed, (2) the currents associated with p- and d − p − 1-form symmetries with mixed anomaly have a defi- nite mathematical relationship, and (3) there is close connection between mixed anomalies, non-equilibrium anomalies, SSB, and propagative versus diffusive dispersion relations.

3.2 Higher-form symmetries: a review

Suppose a relativistic quantum field theory in 3 + 1 dimensions has conserved current

µ µ J , ∂µJ = 0. (3.1)

The usual way to define a conserved charge is by integrating the charge density over the volume of space, that is ¹ d Q = d3xJ0, Q = 0, (3.2) dt but there is another perspective that involves differential forms. In particular, consider the Hodge star dual of the current ρ ? Jµνλ = µνλρJ , (3.3) where µνλρ is the Levi-Civita tensor. Notice that ?J is a three-form, which can naturally be integrated over a three-dimensional hyper-surface in spacetime. Let Σ be such a hyper-

75 surface. Then we can define the charge as a function of the hyper-surface by

¹ Q[Σ] = ?J. (3.4) Σ

If we take Σ to be a spatial volume at fixed time, then we recover the original expression for the charge (3.2). Thus the differential form definition of charge contains within it the standard formulation. Notice that the conservation of J implies that d ? J = 0, that is, ?J is closed. Therefore, if we continuously deform the hyper-surface Σ → Σ0 in such a way that it leaves the boundary unchanged and never intersects any charged operator insertions, then Q[Σ] = Q[Σ0]. In this sense, the charge’s dependence on Σ is purely topological. Moreover, this topological dependence can be interpreted as a conservation law in the sense that we can translate Σ through time without changing the total charge as long as we do not encounter any sources (i.e. charged operator insertions).2 Finally, we should comment on the dimensionality of the objects that are conserved. The charge Q[Σ] counts the total number of point-particles on the slice Σ. From another perspective, however, point-particles travel through spacetime along their respective worldlines, so they ought to be thought of as one-dimensional objects

(or D0-branes). Thus the charge Q[Σ] counts the number of worldlines that intersect the hyper-surface Σ. At this point, we can generalize the notion of current and charge. In the interest of complete generality, we will now work in a d + 1 dimensional spacetime. Suppose a current with p + 1 indices

µ0...µp µ0...µp J , ∂µ0 J = 0, (3.5)

that is antisymmetric under interchange of any adjacent pair of indices. Then the Hodge star dual is

ν0...νp ? Jµ1...µd−p = µ1...µd−p ν0...νp J . (3.6)

2Translating Σ through time will generally change the boundary; however, if Σ is a constant time slice, then it extends to spatial infinity and hence has no boundary.

76 Since ?J is a d − p-form, it is naturally integrated over an d − p-dimensional hyper-surface

Σd−p. We therefore define the charge by

¹ Q[Σd−p] = ?J . (3.7) Σd−p

Notice that d ? J = 0, meaning that the dependence of Q on the choice of hyper-surface

Σd−p is topological, which in turn implies that Q is conserved. Finally, since we are integrat- ing ?J over a surface of dimension d − p, the dimension of the conserved objects will not in

general be point-particles traveling along worldlines (unless p = 0). Instead, Q[Σd−p] counts

the number of charged p + 1-dimension objects—or Dp-branes—that intersect Σd−p. We say that a system with conserved current J has an p-form symmetry. Thus, in this language, ordinary symmetries are zero-form symmetries.

We have made reference to charged operator insertions, but so far have not identified what they are. The objects charged under a p-form symmetry must themselves be p-dimensional objects that cannot be composed of lower-dimensional operators. Suppose we have a p-form field ϕp that enjoys some sort of gauge symmetry. For the sake of concreteness, suppose that we have the gauge symmetry ϕp → ϕp + dγ for p − 1-form γ. Then, given a closed

p-dimensional manifold Cp, we may construct the p-dimensional generalization of a Wilson loop by ∫ p i C ϕ W[Cp] = e p . (3.8)

The charge Q[Σd−p] can detect the presence of the p-dimensional Wilson surfaces if and only

if Σd−p and Cp are linked. In this way, Q[Σd−p] counts the number of closed p-dimensional

Wilson surfaces that link with Σd−p.

77 3.2.1 Superfluids at zero temperature

A relativistic superfluid is a system that, in addition to Poincaré symmetry, enjoys a spontaneously broken U(1) internal symmetry N, which corresponds to particle number conservation. Further, superfluids exist at finite charge density. We assume that in the deep infrared, the only gapless mode is the Goldstone ψ associated with the broken charge N. The leading order-action is therefore

¹ q 4 µ S = d xP(X), X = −∂µψ∂ ψ, (3.9) and the current corresponding to N is

0 µ P (X) J = − ∂ µψ. (3.10) U(1) X

µ ∝ µ ∝ Notice that in equilibrium, since charge density is non-zero, JU(1) δ0 , meaning that ∂µψ 0 δµ. Thus, ψ has a time-dependent background hψi = µ0t. It turns out that µ0 can be interpreted as the equilibrium chemical potential. By virtue of possessing a U(1) Goldstone, superfluids also enjoy a two-form symmetry with corresponding current µνλ µνλρ K =  ∂ρψ. (3.11)

µνλ Notice that ∂µK = 0 identically as partial derivatives commute. The fact that K is conserved as a mathematical identity may give some readers pause: is there any physical content to this conservation equation? It turns out the answer is emphatically, “yes!” To see how this can be, we will recast the theory of superfluids in terms of a dual theory involving a two-form gauge field Aµν.

Begin by replacing ∂µψ → Vµ, for some arbitrary one-form V. Then we can define an auxiliary action ¹   4 1 µνλρ SAUX = d x P(V) −  Vµ∂ν Aλρ , (3.12) 2

78 where V is the magnitude of Vµ. Notice that the equations of motion for Aµν are

∂[µVν] = 0, (3.13)

which, if spacetime is topologically trivial, implies Vµ = ∂µψ for some scalar ψ. We therefore recover the original superfluid action (3.9). Alternatively, we can integrate out Vµ, in which case we find the equations of motion

P0(V) − V µ = F µ, (3.14) V where we have defined the field strength by

µ 1 µνλρ F =  ∂ν Aλρ. (3.15) 2

We can then algebraically solve (3.14) to find V in terms of F.3 Plugging this result back

p µ into (3.12), we obtain a dual action that only depends on Y = −F Fµ. Explicitly, we have

¹ 4 SDUAL = d xL(Y), (3.16) for some function L, where L(Y) ≡ P(V(Y)). Notice that this action is invariant under the gauge symmetry

Aµν → Aµν + ∂[µλν](x). (3.17)

The equations of motion of the dual action are

∂SDUAL ∂µ = 0. (3.18) ∂(dA)µνλ

3In general the equation to be solved may be very complicated, but all we require is that a solution exist.

79 But this equation is just the conservation law for a higher-form current. We therefore identify

∂SDUAL K µνλ = . (3.19) ∂(dA)µνλ

We now see that the conservation of K is no longer a mere mathematical identity, but is a manifestly physical equation describing the dynamics of superfluids. But what has happened µ to JU(1)? Comparing equations (3.10) and (3.14), we have the identification

µ µ JU(1) = F . (3.20)

µ But notice that ∂µF = 0 identically owing to the fact that partial derivatives commute. Thus, in the dual picture the particle-number conservation equation is a mere mathematical identity.

3.2.2 Electromagnetism

We now consider the theory of electromagnetism as a system involving higher-form sym- metries. Letting A be the photon field and F = dA be the field-strength tensor, the action for free Maxwell theory in 3+1 dimensions is

1 ¹ S = − d4xF2. (3.21) 4

µν The equations of motion are ∂µF = 0, meaning that we have a one-form symmetry with µν ≡ µν corresponding conserved current Jel F . Then given a closed two-dimensional surface

Σ2, we have the conserved charge

¹ Q[Σ2] = ?F. (3.22) Σ2

80 Taking Σ2 to exist at constant time, we therefore have

¹ Q[Σ2] = nˆ · Eì, (3.23) Σ2

i 0i where nˆ is normal to the surface Σ2 and E = F . We thus recognize this one-form charge

as counting the number of electric field lines passing through the Gaussian surface Σ2. What are the charged objects? As we mentioned earlier, they are Wilson loops. For closed curve C, we have i ∫ A W[C] = e C . (3.24)

Physically, C corresponds to the worldline of a massive charged particle which is a source

for the electric field lines. In this way, if C links with Σ2, then the amount of electric flux

through Σ2 changes according to Gauss’s law. We can see, therefore, that Q[Σ2] counts the

number of Wilson loops that link with Σ2. µν Electromagnetism has another one-form symmetry with corresponding current Jmag = ?F µν, which is conserved as an identity. This charge counts the number of magnetic field lines passing through a given Gaussian surface. When dealing with the full, sourced Maxwell µν theory, Jel is no longer exactly conserved, whereas, in the absence of magnetic monopoles, µν Jmag is always exactly conserved. As a result, when we consider magnetohydrodynamics, µν Jmag will play a key role.

3.3 Non-equilibrium cosets and the classical limit

We will review some of the key points regarding the non-equilibrium coset construction developed in the previous chapter and then show explicitly how the classical limit may be taken. Suppose a d + 1-dimensional relativistic system with global (zero-form) symmetry group G, which includes both internal and spacetime symmetries and is spontaneously broken to

81 the subgroup H. We denote the generators by

P¯µ = unbroken translations,

TA = other unbroken generators, (3.25)

τα = broken generators.

The unbroken subgroup H is then generated by P¯µ and TA. We assume the existence of d + 1 unbroken translation generators P¯µ, but do not require them to be the spacetime translation generators of the Poincaré group. We construct the effective action on the ‘fluid worldvolume’ coordinates σM for M = 0,..., d and parameterize the most general elements (one for each leg of the SK contour) of G by

µ α A iX (σ)P¯µ iπ (σ)τα ib (σ)TA gs(σ) = e s e s e s , s = 1, 2. (3.26)

Under a global symmetry transformation γ ∈ G, transformations of the Goldstone fields can be computed according to

0 0 0 gs[Xs, πs, bs] → gs[Xs, πs, bs] ≡ γ · gs[Xs, πs, bs]. (3.27)

In order to distinguish the broken and unbroken symmetries, we require that the EFT be invariant under certain time-independent gauge transformations.4 Throughout the rest of the chapter, we let indices M, N, P, Q = 0,..., d be coordinate indices of σ and we let I, J, K, L = 1,..., d be spatial coordinate indices of σ. First, since we are assuming the existence of d + 1 translation generators, impose the time-independent diffeomorphism symmetry on the coordinates σ0 → σ0 + f (σI), σI → ΣI(σJ), (3.28)

for arbitrary f and ΣI. If there were fewer than d + 1 unbroken translations, then this set of 4We say that these transformations are gauge symmetries because they do not change the state of the system, but are instead redundancies of description.

82 coordinate symmetries could be reduced; see [32]. Next, because TA are unbroken, we have the right-acting, time-independent gauge symmetries

A I iλ (σ )TA gs(σ) → gs(σ)· e . (3.29)

Finally, we compute the Maurer-Cartan form, which is a Lie algebra-valued one-form −1 given by gs dgs. Because it is Lie algebra-valued, it is always expressible as a linear combi- nation of symmetry generators. In particular, we have

−1 M  µ ¯ α  A  gs dgs = idσ EsM Pµ + i∇µπs τα + BsMTA . (3.30)

µ It can be checked that under the gauge symmetries (3.28) and (3.29), EsM transform as α A vielbeins, ∇sµπ transform linearly and hence we call them ‘covariant derivatives,’ and BsM transform as gauge fields. The effective action is then constructed by forming manifestly symmetry-invariant terms out of the vierbeins, covariant derivatives, and connections. Finally, suppose that Θ is some anti-unitary, time-reversing symmetry of the ultraviolet theory and that the equilibrium density matrix describes a state in thermal equilibrium, namely ¯0 e−β0P ρ = ¯0 , (3.31) tre−β0P where β0 is the equilibrium inverse temperature. Then, the effective action must be invariant under the so-called dynamical KMS symmetries, which are non-local Z2 symmetries whose actions are5

θP¯0 0 g1(σ) → Θe g1(σ − iθ, σì ), (3.32) −(β0−θ)P¯0 0 g2(σ) → Θe g2(σ + i(β0 − θ), σì ),

for arbitrary θ ∈ [0, β0] [2]. In the classical limit, these symmetries reduce to a single local

5 θ P¯0 −(β0−θ)P¯0 0 The factors e and e arise because the field Xs (σ) transform as a time coordinates under dynamical KMS symmetries.

83 Z2 transformation. To take the classical limit, it is convenient to work in the retarded-advanced basis. In

particular for a given set of fields ϕs(σ), s = 1, 2 defined on the SK contour, we define the retarded-advanced fileds by

1 ϕr = (ϕ1 + ϕ2), ϕa = ϕ1 − ϕ2. (3.33) 2

It will be useful to define the a-derivative denoted by δa whose action is given by δaϕr = ϕa, 2 for any retarded-advanced pair of fields ϕr and ϕa. We also require that δa = 0 and that it satisfy the Leibnitz rule, namely

0 0 0 δa(ϕr ϕr ) = ϕaϕr + ϕr ϕa. (3.34)

Then, the way to construct classical invariant building-blocks is as follows. Parameterize the most general group element with retarded Goldstones

µ α A iX (σ)P¯µ iπ (σ)τα ib (σ)TA gr (σ) = e r e r e r , (3.35)

−1 and construct the Maurer-Cartan form gr dgr , which is given by (3.30) with replacement s → r. Then, we can construct retarded building-blocks from this Maurer-Cartan form in the usual way. To construct advanced building-blocks, we need only act with δa on terms of the retarded Maurer-Cartan form. The classical dynamical KMS symmetry is then

∂gr (σ) gr (σ) → Θgr (σ), ga(σ) → Θga(σ) − iΘβ0 + Θβ0P¯0gr (σ), (3.36) ∂σ0

where ga ≡ δagr .

84 3.4 Non-equilibrium ’t Hooft anomalies

In non-equilibrium systems, it is possible to have a symmetry with no corresponding non-trivial Noether current. As a simple example, consider a non-relativistic point particle in a uniform fluid undergoing Brownian motion. The effective action for the point particle is given by ¹  i 2 Ip.p. = dt − νxa xÛr + M˜ xÛr xÛa + σx , (3.37) 2 a

where the retarded field xr (t) gives the classical position of the particle, xa(t) is the corre- 1 sponding advanced field, and the dynamical KMS symmetry requires that ν = 2 σβ0 ≥ 0, which is just a statement of the fluctuation-dissipation theorem. Evidently, the physics of the point particle is independent of its position in space and time as the action is invariant under

t → t + c0, xr → xr + c1, (3.38)

for constants c0 and c1. In other words, it enjoys spacetime translation symmetry. Given that it exists in a medium, however, the particle is subject to friction and hence its energy and momentum are not conserved. In particular, the equations of motion are

M˜ xÜr = −νxÛr . (3.39)

Thus, the standard particle momentum associated with spatial translation symmetry is not conserved.6 This finding may puzzle some readers: we have an action with a global symmetry, meaning that the assumptions of Noether’s theorem are satisfied. So how can there be no conserved current? The answer is that at the level of the mathematics, there is a conserved current associated with every symmetry, but it need not contain any physical information.

6 Notice that there is in fact a conserved quantity here, namely M˜ xÛr +νxr . That such a conserved quantity exists, however, is accidental. If we were to include higher-order terms in the Lagrangian like −xa f (xÛr ) + ... , then the equations of motion would read M˜ xÜr = −νxÛr − f (xÛr ), which admits no conserved quantities for generic function f . Thus, the most general EFT yields no conserved momentum despite enjoying exact spatial translation symmetry.

85 To see how this is so, let us compute the Noether current associated with spatial translations. We have7

j = −νxa + M˜ xÛa. (3.40)

But when the equations of motion are satisfied, all advanced fields vanish, so we have xa = 0 and as a result j = 0. Thus the conservation equation ∂t j = 0, while true, provides no physical information about the system. At the level of non-equilibrium EFTs, the existence of a conserved current is closely related to gaugeability. In particular, if a given global symmetry can be gauged by introducing two copies—one for each leg of the SK contour—of a gauge field, then there exists a corresponding conserved current. For further explanation, see §3.6.1 or consult [31, 32, 30, 29]. We refer to such anomalous charges as non-equilibrium ’t Hooft anomalies because they arise when a symmetry of a non-equilibrium system cannot be gauged. In the point-particle example, the fact that the energy and momentum exhibit a non- equilibrium ’t Hoof anomaly owes to the fact that the particle is an open system and can freely exchange energy and momentum with an environment. Such anomalous behavior, however, does not always arise because of interactions with an environment. For example, the diffusive phason mode in a quasicrystal owes its existence to a U(1) shift symmetry with no corresponding Noether current [31]. We will investigate this example and will see other examples of non-equilibrium ’t Hooft anomalies that arise in closed systems in subsequent sections of this chapter. It can be shown that if we build the the effective action exclusively from terms furnished by the coset construction, then our Goldstone fields are gaugeable; see §3.6.1. This obser- vation raises the questions: what kinds of terms are invariant under the global symmetry group but are not gaugeable? To answer this question, consider

−1 ga(σ) ≡ g2 (σ)· g1(σ). (3.41) 7Notice that j has no indices because we are working with a 0 + 1-dimensional QFT.

86 Evidently ga transforms by conjugation under the gauge symmetry defined by (3.29) and is

invariant under all global (i.e. physical) G-symmetry transformations. Notice that ga ∈ G,

while the Maurer-Cartan forms are Lie-algebra valued. To turn ga into a Lie algebra-valued object, we need only take the log, that is

Ga ≡ log ga. (3.42)

We can then expand Ga as a linear combination of symmetry generators

= iΦµ P¯ + iΦατ + iΦAT . Ga P¯ µ τ α T A (3.43)

If a given symmetry exhibits a non-equilibrium ’t Hooft anomaly, we must include the cor- responding Φ as a building-block in the effective action. For example if we want τα to be a

α symmetry with no corresponding conserved current, then we need only use Φτ as a covariant building-block in the EFT. −1 Taking the classical limit, we have Ga = gr δagr . Thus, Ga becomes like a Maurer-Cartan form with respect to the “derivative” defined by δa.

3.4.1 Reactive fluids

We will now demonstrate how to use the coset construction to formulate an EFT for fluids that exhibit reactive flows. There are, in addition to Poincaré symmetry, n different unbroken U(1) charges that we denote by N1,..., Nn. Each U(1) charge counts the number of particles of a given species. To allow for chemical reactions, that is, the transformations of certain types of particles into other types of particles, we must kill the conservation of some subset of these charges. But first, we will construct the action for which there are no chemical reactions. It turns out that we will need to use the full SK contour to describe these dynamics, so we must construct an action with doubled field content. We will work in the classical limit,

87 so we can use the δa trick to simplify matters. The most general retarded group element is

µ A i i iX (σ)Pµ iψ (σ)NA iη (σ)Ki iθ (σ)Ji gr (σ) = e r e r e r e r , (3.44)

where A = 1,..., N. Because translations are unbroken, we impose the fluid diffeomorphism

A symmetry (3.28) and because Ji and N are unbroken we have the gauge symmetries

i I A I iλ (σ )Ji ic (σ )NA gr (σ) → gr (σ)· e , gr (σ) → gr (σ)· e . (3.45)

The resulting retarded Maurer-Cartan form is then

−1 µ i i A gr dgr = iEr (Pµ + ∇µηr Ki) + iωr Ji + iBr NA, (3.46)

where

µ ν µ ErM = ∂M Xr [Λr Rr ]ν ,

i −1 M −1 0j ji ∇µηr = (Er )µ [Λr ∂M Λr ] Rr , 1 (3.47) ωi = i jk[(Λ R )−1∂ (Λ R )]jk, rM 2 r r M r r a BrM = ∂M ψr,

i i µ iη Ki µ ij iθ Ji ij such that Λr ν = (e r ) ν and Rr = (e r ) . To remove Lorentz Goldstones, we impose the inverse Higgs constraints

i i ηr Er0 = 0 =⇒ tanh ηr, (3.48) ηr

and i jk −1 M k i jk −1 M k  (Er )j EaM = 0,  (Er )j ∂0ErM = 0, (3.49)

µ µ where EaM = δaErM . This second inverse Higgs constraint can be solved to remove the

88 rotation Goldstones, but it is not necessary at this level in the derivative expansion; see [29]. After imposing these inverse Higgs constraints, we are left with the invariant building- blocks A 1 A 1 ∂ψr T = √ , µ = √ 0 , (3.50) −Gr00 −Gr00 ∂σ and the invariant integration measure is

4 p d σ −Gr . (3.51)

µ ν As before, we have defined the fluid metrics by GrMN = ErM ηµν ErN .

To construct advanced building-blocks, we need only act with δa on retarded building- blocks. At leading order, the dynamical KMS symmetries allow us to write

IEFT = δaS, (3.52) where ¹ 4 p A S = d σ −Gr P(T, µ ). (3.53)

µ µ At this point, it is convenient to transform to the physical spacetime, so letting x ≡ Xr , we have8 ¹ S = d4xP(T, µA). (3.54)

To be explicit about the form of the non-equilibrium effective action, we may write

¹ 4  µν Aµ A I = d x T ∂µXaν + J ∂µψ , (3.55) where the stress-energy tensor is

T µν = (T, µA)uµuν + p(T, µA)∆µν, (3.56)

8The fluid coordinates are now dynamical fields σM (x).

89 such that ∆µν = ηµν + uµuν and the U(1) currents associated with N A are

J Aµ = nA(T, µB)uµ. (3.57)

The (classical) dynamical KMS symmetry imposes the relations

∂p ∂p  + p = T + µA , ∂T ∂µA (3.58) ∂p nA = , ∂µA

which are equivalent to the local first law of thermodynamics. The equations of motion are just the conservation of energy, momentum and charge, namely

µν Aµ ∂µT = 0, ∂µJ = 0. (3.59)

Since J Aµ are all independently conserved, we see that our action describes fluids with n species of particles that are not permitted to undergo any chemical reactions. To include chemical reactions, we must kill certain linear combinations of U(1) conser- vation equations, while preserving other linear combinations. Such non-conservation can be accomplished by introducing non-equilibrium ’t Hooft anomalies for certain charges. Without ˆ loss of generality, let N A for Aˆ = 1,..., k < n be representatives of the coset of non-conserved U(1) charges. Then we have invariant building-blocks furnished by

−1 gr δagr ⊃ iψAˆNAˆ. (3.60)

We therefore find that new terms can be added to (3.55) and obtain

¹   4 µν Aµ A Aˆ Aˆ i AˆBˆ Aˆ Bˆ I = d x T ∂µXaν + J ∂µψ − Γ ψ + M ψ ψ , (3.61) a 2 a a

90 where the dynamical KMS symmetry imposes

Aˆ 1 AˆBˆ µ Bˆ Γ = M β ∂µψ . (3.62) 2 r

µ The equations of motion for Xa are still the conservation of the stress-energy tensor. But A the equations for ψa are now Aµ A ∂µJ = −Γ , (3.63)

ˆ ˆ where ΓA ≡ δAAΓA. But this is just the equation for k non-conserved U(1) currents and n − k conserved U(1) currents. Thus, chemical reactions may take place. These results agree with those of [30].

3.4.2 Second sound and its removal

In condensed matter systems at finite temperature, it is sometimes the case that there are two longitudinal sound modes: one of which is the Goldstone associated with some spon- taneously broken spacetime translation symmetry while the other is the usual hydrodynamic sound mode associated with conservation of the stress-energy tensor. The most commonly studied second sound modes emerge in superfluids. At zero temperature, superfluids enjoy a single sound mode associated with the spontaneous breaking of time-translation symme- try. When the temperature becomes non-zero, a weakly-coupled gas of superfluid phonons emerges. In this gas of phonons, collective hydrodynamic waves may propagate with a differ- ent speed than the ordinary superfluid sound mode. In this way, superfluids have a second, hydrodynamic sound mode. Similar such second sound modes have also been observed in solids; however unlike superfluids, solid second sound modes require very particular condi- tions to exist. In the vast majority of solids, no second sound can be observed. For second sound to exist, the solids must enjoy a pristine lattice structure and must be cooled to suffi- ciently low temperatures that Umklapp scattering is negligible [59]. At the level of our solid EFT, these conditions are equivalent to having conserved lattice momentum—Umklapp scat-

91 tering is the process by which lattice momentum is not conserved. The generators of lattice

momentum are Qi for i = 1, 2, 3, introduced in § 1.3. In the previous chapter, we constructed actions for solids and smectic liquid crystals all with second sound. In this subsection, we will see how second sound can be removed. Let us begin by investigating the solid action; the smectic actions will follow in similar fashions. The internal translation Goldstones are ψi for i = 1, 2, 3. We wish to destroy the conservation of lattice momentum, while keeping the notion of large-scale isotropy. Thus,

we want Qi to represent exact symmetries of the action, yet have no associated conserved

current. We therefore must endow Qi with non-equilibrium ’t Hooft anomalies. It turns out that even at leading-order in the derivative expansion, if we wish to give lattice momentum non-equilibrium ’t Hooft anomalies, we must work with the full non-equilibrium effective action, including the double field content. Consider the solid action (2.57); to formulate the

action with doubled field content, we need only apply δa to SEFT, that is IEFT = δaSEFT. Explicitly, we have ¹   4 1 µν iµ i IEFT = d x T Gaµν + J C , (3.64) 2 aµ

µν iµ i i where T and J are given by (2.61) and (2.64), Gaµν = ∂µXaν + ∂ν Xaµ, Caµ = ∂µψa, and we have performed a coordinate change so that our action is defined on the physical spacetime

µ µ coordinates x ≡ Xr . Now that our action is defined on the physical spacetime, the fluid worldvolume coordinates become dynamical fields, namely σM (x).

i If Qi have non-equilibrium ’t Hooft anomalies, then ψa are allowable building-blocks. Thus, our action becomes

¹   0 4 1 µν iµ i i i i ij i j I = d x T Gaµν + J C − Γ ψ + M ψ ψ , (3.65) EFT 2 aµ a 2 a a

where the dynamical KMS symmetries impose the relationship

i β0 ∂ψ Γi = Mi j r . (3.66) 2 ∂σ0

92 The equations of motion for ψi are now

iµ i ∂µJ = Γ ; (3.67)

we see that Qi are no longer conserved. Notice that in the deep infrared, the l.h.s. of the above equation becomes negligible as it is higher-order in the derivative expansion. As a result, if we are working only to leading-order in derivatives, the equations of motion for ψi become Γi = 0. (3.68)

But the dynamical KMS relation (3.66) can be used to reduce this equation to

∂ψi r = 0. (3.69) ∂σ0

i Also, at leading order in the derivative expansion, we have ψa = 0. Recall from the previous chapter that we had an expanded definition of inverse Higgs constrains: inverse Higgs con- straints arise whenever symmetry-invariant terms can be set to zero such that when solved, one set of fields can be removed in favor of another. We claim that the equations

∂ψi r = 0, ψi = 0, (3.70) ∂σ0 a satisfy this definition of inverse Higgs constrains. To see that this is so, notice that the first

i I equation requires that ψr be functions of the spatial fluid worldvolume coordinates σ for I = 1, 2, 3. As a result, we may write

i i ψr = F (σì ), (3.71) for some σ0-independent functions Fi. But notice that the fluid worldvolume coordinates σi have full spatial diffeomorphism gauge symmetry. We may therefore choose the gauge-fixing

93 conditions i i ψr = σ . (3.72)

i Thus, we have fully removed the fields ψr at the price of restricting the diffeomorphism sym- metry of the fluid worldvome. We call this partially gauge-fixed space the ‘solid worldvolume’ and it enjoys the residual diffeomorphism symmetries.

σ0 → σ0 + f (σì ), σI → σI + cI, (3.73) for arbitrary time-independent function f and constants cI. If the solid is isotropic then the worldvolume coordinates enjoy the additional internal rotation symmetry σI → RIJ σJ, for R ∈ SO(3). Analogous procedures may be implemented to obtain the Smectic A and B worldvolumes, whose gauge symmetries are given respectively by

σ0 → σ0 + f (σì ), σA → σA + gA(σì ), σ3 → σ3 + c3, (3.74) and σ0 → σ0 + f (σì ), σA → σA + gA(σ3), σ3 → σ3 + c3, (3.75) where we use index A = 1, 2 to indicate directions perpendicular to the 3-direction.

3.4.3 Dissipative solids and smectics without second sound

We begin by constructing the effective action for solids without second sound to leading order in dissipation; then we will show how to modify it to account for smectics in phases A and B. The construction of such an action requires doubled field content. We could use the coset construction to formulate this higher-order action, but we find it convenient to use a different method that makes the constitutive relations more apparent. To make our model of solids more physically realistic, we will suppose an unbroken U(1) charge N, corresponding to particle-number conservation.

94 We will work exclusively in the classical limit, allowing us to formulate our action on

µ µ the physical spacetime coordinate x ≡ Xr . As a result, the solid worldvolume coordinates M µ become dynamical fields. Our field content is now σ (x), ϕr (x), Xa (x), and ϕa(x), where

ϕr/a are the Stückleberg fields associated with N. To construct the effective action, it is helpful to first identify the symmetry-covariant

M M building-blocks. The retarded building-blocks are as follows. Let eµ (x) ≡ ∂µσ (x). Then the local inverse-temperature four-vector field is given by

µ −1 µ β (x) = β0(e )0, (3.76)

where β0 is the equilibrium inverse temperature. This field encodes all information about the temperature and local rest-frame of the solid volume elements. It is often helpful to decompose this object into its magnitude and direction by βµ = βuµ, where

p βµ β = −β2, uµ = . (3.77) β

Next, there are the solid basis vectors and (inverse) solid metric given respectively by

I IJ I µν J eµ(x), γ (x) = eµη eν . (3.78)

Note that I, J, K, L = 1, 2, 3 indicate spatial solid coordinate indices. We also have the chem- ical potential µ given by µ µ = u ∂µϕr, (3.79)

as an invariant building-block. In addition to the retarded building-blocks, we have the advanced covariant building-blocks

Gaµν = ∂µXaν + ∂ν Xaµ, Baµ = ∂µϕa. (3.80)

95 µ IJ The leading order action is constructed with β, µ, u , γ , Gaµν, and Baµ without any additional derivatives. Further, because advanced fields count at higher-order in the deriva- tive expansion, the leading terms may only contain one factor of Gaµν and Baµ. Thus, the leading-order Lagrangian is 1 µν µ L1 = T Gaµν + J Baµ, (3.81) 2 0 0

µν µ µ where T0 is some symmetric tensor and J0 is some four-vector built from β, µ, u , and γIJ and have the interpretations of the stress-energy tensor and particle number current, µ µν respectively. Notice that the equations of motion for Xa and ϕa are, respectively ∂νT0 = 0 µ and ∂µJ0 = 0. The most general form this leading-order stress-energy tensor can take is

µν µ ν µν IJ µν T0 = 0u u + p0∆ + rIJ∆ , (3.82)

µν µν µ ν IJ 1 I J J I where ∆ = g + u u and ∆µν = 2 (eµeν + eµeν). And the leading-order particle number current is µ µ J0 = n0u . (3.83)

IJ We take 0, p0, rIJ, and n0 to be generic functions of β, µ, and γ . For isotropic solids, the sum over indices I, J must be performed in a rotationally-invariant manner, but for generic

IJ solids, the I, J indices on rIJ and γ are purely for the purposes of bookkeeping and need not transform in any particular way. At leading order, the dynamical KMS symmetries allow us to write the action in factorized ∫ 4 3 form as d xL1 = SEFT[X1] − SEFT[X2] + O(a ), where SEFT is an ordinary action that only

depends on one copy of the fields. As a result, we have the relations among 0, p0 and rIJ given by

IJ ∂p0 ∂p0 ∂p0 ∂p0 p0 ≡ p0(β, µ, γ ), 0 + p0 = −β + µ , rIJ = , n0 = . (3.84) ∂β ∂µ ∂γIJ ∂µ

The next-to-leading-order (NLO) terms in the Lagrangian give rise to dissipation. We

96 have 1 µν µ i µν,αβ µν L2 = T + J Baµ + W GaµνGaαβ + iZ BaµBaν. (3.85) 2 1 1 4 0 0

µν µ Physically, T1 and J1 are, respectively the NLO contributions to the stress-energy tensor µν,αβ µν and the particle number current. By contrast, W0 and Z0 encode information about statistical fluctuations. The explicit forms of these terms are potentially quite complicated. Fortunately, we can employ the generalized Landau frame [2] to simplify matters. In partic- ular, we have

µν IJ µν KLαβ µν,αβ −1 IJ µν KLαβ T1 = −ηIJKL∆ ∆ ∂αuβ, W0 = β ηIJKL∆ ∆ , (3.86)

and µ IJ µν −1 µν −1 IJ µν J1 = −σIJ∆ β ∂ν(βµ), Z0 = β σIJ∆ , (3.87)

µν,αβ µν where the forms of W0 and Z0 are determined by the classical dynamical KMS symme- 9 tries. We interpret ηIJKL as the viscosity tensor and σIJ as the charge conductivity tensor,

which is related to the thermal conductivity tensor κIJ by

 2 0 + p0 κIJ = βσIJ . (3.88) n0

These tensors are generic functions of β, µ, and γIJ and enjoy various symmetries among their indices, namely

ηIJKL = ηJIKL = ηIJLK = ηKLIJ, σIJ = σJI . (3.89)

Putting it all together, the full Lagrangian to leading order in dissipation is L = L1 + L2, 9In general, the field redefinitions required to arrive at the generalized Landau frame do not respect the dynamical KMS symmetries. Conveniently, however, the dynamical KMS symmetries hold for the NLO Lagrangian in generalized Landau frame.

97 or explicitly,

1   i  L =  uµuν + p ∆µν + r ∆IJ µν − η ∆IJ µν∆KLαβ ∂ u − G G 2 0 0 IJ IJKL α β 4β aαβ aµν (3.90)  µ IJ µν −1  + n0u − σIJ∆ β ∂ν(βµ) − iBaν Baµ.

Lastly, it is worth noting that the effective actions for smectic liquid crystals in phases A and B can be obtained quite easily from the above action. In fact, they are special cases of the above. Consider the symmetries (3.74) and (3.75). Notice that they contain strictly more symmetries than the (anisotropic) solid worldvolume. In this way smectic liquid crystals can be seen as symmetry-enhanced solids. To obtain the action for smectic A liquid crystals, we restrict the kind of dependence the action can have on γIJ. In particular, the action 33 may only depend on γ and the transport coefficients ηIJKL and σIJ must be rotationally symmetric about the 3 direction. Explicitly, for any rotation

IJ cos − sin 0 © θ θ ª ­ ® IJ( ) ­ ® R θ = ­sin θ cos θ 0® , (3.91) ­ ® ­ 0 0 1® « ¬

we have that

II 0 JJ0 KK 0 LL0 II 0 JJ0 ηIJKL = R R R R ηI 0J0K 0L0, σIJ = R R σI 0J0 . (3.92)

The action for smectic B liquid crystals has the same constraints as that of phase A except its dependence on γIJ may also include

11 33 13 2 22 33 23 2 b = det γ, b1 = γ γ − (γ ) , b2 = γ γ − (γ ) , (3.93)

and ηIJKL need not enjoy any particular rotational symmetry. Comparing with the results of §1.6.3, we see that fluids and smectic liquid crystals can be viewed as highly symmetric

98 symmetric solids at the level of the non-equilibrium effective action. Recent papers [77, 57] have construct effective field theories for relativistic solids by for- mulating constitutive relations for conserved quantities. The constitutive relations that these authors arrive at bear great resemblance to those presented here, but with some differences.

I In particular, in our formulation, the solid basis vectors eµ are automatically orthogonal to µ µ I the fluid velocity u . That is u eµ ≡ 0 off shell. This orthogonality arises because we have removed second sound; if second sound were present then the fluid degrees of freedom could

µ I flow freely relative to the solid degrees of freedom, thereby permitting u eµ , 0. Alter- natively even with second sound present, we could abstain from imposing the ψs-removing

µ I inverse Higgs constraints and find that, on the equations of u eµ is non-zero but decays to zero exponentially fast. In particular if ΓI is sufficiently large, then the exponential decay will take place on shorter time scales than the UV cutoff of the EFT; if ΓI is sufficiently small, then such decays occur on time scales longer than the UV cutoff. As a result, the results presented in this section agree with those of [77, 57] so long as

µ I we augment their equations with additional equations either forcing u eµ ≡ 0 or equations dictating the (non)-conservation of the lattice momentum currents JI µ.

3.4.4 Quasicrystals

Quasicrystals, or quasi-periodic structures, are materials with perfect long-range order that lack periodicity. As an example, consider the one-dimensional function

k1 f (x) = cos k1x + cos k2x, < Q. (3.94) k2

Evidently, there is long-range order because f is the sum of two periodic functions, but the periods of the cosines are not rational multiples of each other, so f will never repeat. More generally, a quasi-periodic structure in d dimensions can be formed by first consider- ing a periodic lattice structure in a higher dimensional space with dimension D > d. Next,

99 consider a d-dimensional sub-manifold that intersects the D-dimensional lattice at an incom- mensurate angle with respect to at least one lattice vector. Then the resulting lattice-like structure inherited by the d-dimensional sub-manifold will exhibit long-range order owing to the fact that the D-dimensional lattice is periodic, but the pattern will never repeat, owing to the fact that the angle is incommensurate. If we translate the d-dimensional sub- manifold in an orthogonal direction, then the quasi-periodic structure changes. Suppose the quasi-periodic structure represents the atomic cites of a solid-like substance. The free-energy of a quasicrystal must remain unchanged after such atomic rearrangements associated with orthogonal translations. We will use the intuition provided by this embedding picture to help facilitate the con- struction of the quasicrystal EFT. In particular, we will embed the physical three-dimensional

A space into a four-dimensional space with coordinates ψs for A = 1, 2, 3, 4 and s = 1, 2, as shown in figure 3.1.10 If we neglected the A = 4 field, then our construction would look just

i like that of an ordinary solid, with the fields ψs for i = 1, 2, 3 being the ordinary solid fields 4 generated by Qi. The remaining filed, ψs is a new field that translates the embedded space i in the orthogonal direction; it is generated by Q4. Physically, ψs are the phonons—they de- 4 scribe pressure wave through the quasicrystal; ψs is known as the phason field and describes 4 4 4 microscopic rearrangements of the atoms. A constant shift ψs → ψs + c will rearrange all of the atoms, but will leave the free-energy unchanged. We therefore postulate that Q4 generates a symmetry of our effective action. But since Q4 is not a fundamental symmetry of the UV physics, we expect no corresponding Noether current to exist. We therefore must endow Q4 with a non-equilibrium ’t Hooft anomaly. We will additionally suppose that no second sound exists, meaning that Qi also exhibit non-equilibrium ’t Hooft anomalies.

The internal shift symmetries generated by QA are

A A A ψs → ψs + c , (3.95) 10Notice that unlike the fluid worldvolume, the quasicrystal space has no time-like coordinate.

100 Qusicrystal space (2) Fluid (spatial) worldvolume Quasicrystal space (1) A A ψ1 (σ) ψ2 (σ) σi

Figure 3.1: This figure depicts how, at fixed time σ0, the spatial submanifold of the fluid worldvolume is mappeed into two copies of the quasicrystal worldvolume with the embedding A B 0 maps ψ1 (σ) and ψ1 (σ). We suppress the σ -coordinate as the quasicrystal spaces have no intrinsic notion of time. Notice that unlike the mapping into physical spacetime, the quasicrystal dimension is greater than that of the spatial fluid worldvolume. for constants cA. Notice that because our EFT is defined on the SK contour, in the dis-

A A tant future ψ1 (+∞) = ψ2 (+∞). As a result, there can be only one copy of the above shift symmetries despite the fact that the field content is doubled. The coordinates ψ A represent coordinates of a four-dimensional lattice. Such lattices often enjoy certain rotational sym- metries. Suppose these rotational symmetries form the discrete subgroup S ⊂ SO(4), then we expect the EFT to be invariant under the transformations

A AB B ψs → O ψs , O ∈ S. (3.96)

We are now in a position to construct the quasicrystal EFT. Let us begin by constructing covariant terms. For simplicity, we are only interested in the classical EFT. Taking the classical limit is most conveniently performed in the retarded-advanced basis. Thus, we will construct covariant terms using the fields

1 1 X µ ≡ (X µ + X µ), X µ ≡ X µ − X µ , ψ A ≡ (ψ A + ψ A), ψ A = ψ A − ψ A . (3.97) r 2 1 2 a 1 2 r 2 1 2 a 1 2

101 As in the solid case, we may perform a coordinate transformation so that our action is

µ µ defined on the physical spacetime x = Xr ; then the fluid coordinates become dynamical M M M fields σ (x). And once again, let eµ = ∂µσ . Then, we again have the covariant building- blocks µ −1 µ µ β ≡ β0 (e )0 ≡ βu , Gaµν ≡ ∂µXaν + ∂ν Xaµ, (3.98) where β0 is the equilibrium inverse temperature. Additionally, we have covariant building

A blocks constructed using the quasicrystal fields. Notice that (3.95) leaves ψa invariant while A it shifts ψr by a constant. Therefore,

A A ∂µψr , ψa , (3.99) are covariant building-blocks of our EFT. The last consideration before we can construct the leading-order effective action is to determine how the fields transform under the classical dynamical KMS symmetries. The spacetime embedding fields transform by

µ µ µ µ µ µ Xr → ΘXr , Xa → ΘXa − i Θ β + i β0 δ0, (3.100) and the quasicrysal fields transform by

A A A A µ A ψr → Θ ψr , ψa → Θ ψa − i Θ β ∂µψr , (3.101) where Θ is a time-reversing symmetry transformation of the ultraviolet theory. It is conve-

AB A µ B nient to define the symmetric matrix Y ≡ ∂µψr ∂ ψr , and the column vector (in quasicrys- A µ A tal space) Z ≡ β ∂µψr . µ i In the derivative expansion, β and ∂µψ for i = 1, 2, 3 count as zeroth order because their equilibrium expectation values are non-vanishing constants. Therefore the effective action may depend on arbitrary Poincaré-invariant functions of these building-blocks. The phason

102 4 building-block ∂µψr , however, has vanishing expectation value. We therefore count it as first A µ order in derivatives. Similarly, we count ψa and Xa , which all have vanishing expectation A A value, at first order in derivatives. Hence count Cµ = ∂µψa and Gaµν = ∂µXaν + ∂ν Xaν at second order in derivatives. Thus, performing a coordinate transformation so that our EFT

µ µ is defined on the physical spacetime x ≡ Xr , the leading order Lagrangian with non-trivial dynamics is 1 µν Aµ A A A i AB A B LEFT = T Gaµν + J C − Γ ψ + M ψ ψ , (3.102) 2 µ a 2 a a

where

µν AB A µ ν AB A µν AB AB A µ A ν B T = (β,Y , Z ) u u + p(β,Y , Z ) ∆ + r (β,Y , Z ) ∂ ψr ∂ ψr (3.103)

is the stress-energy tensor such that ∆µν = ηµν + uµuν,

Aµ A AB A µ AB AB A µ B J = F (β,Y , Z ) u + H (β,Y , Z ) ∂ ψr , (3.104)

and ΓA and M AB are generic functions of β, Y AB, and Z A. Further, unitarity mandates that M AB be positive definite, which has the additional effect of ensuring the path integral

converges. Notice that non-derivative terms in the ψr field (but not in the ψa one) are forbidden because of the shift symmetry (3.95). The various coefficient functions, , p, r AB, F A, H AB, ΓA, and M AB are not totally in- dependent of one another as we must impose the (classical) dynamical KMS symmetry. It is straightforward (though perhaps tedious) to show that the dynamical KMS symmetry imposes the following relations:

∂p ∂p ∂p ∂p β0  + p = −β , r AB = , F A = , H AB = 2 , ΓA = M AB Z B. (3.105) ∂β ∂Y AB ∂Z A ∂Y AB 2

The first equation is nothing other that the standard thermodynamic law  + p = sT. With

103 the relations (3.105), this Lagrangian (3.102) is the most generic EFT at leading order in the derivative expansion that is consistent with symmetries.

i Following an almost identical procedure as in the solid case, we may remove ψr/a entirely at the price of reduced gauge symmetries on the worldvolume. In particular, we have

σ0 → σ0 + f (σi), σi → σi + ci, (3.106) where f is a generic function of the spatial coordinates and ci are constants. As before, we call this space the solid worldvoume. Now the remaining fields are11

µ µ 4 4 Xr (σ), Xa (σ), ψr (σ), ψa (σ) . (3.107)

The first two fields describe the embedding into the physical spacetime and the last two describe fluctuations in the quasicrystal space perpendicular to the solid worldvolume and hence correspond to the phason degrees of freedom.12

i i i Since we have removed ψr/a, we lose some building blocks, namely Z , and ψa, whereas others are altered, like Y AB. In particular we find that Y AB is replaced by yAB, where13

i j i µν j 4i i4 i µ 4 44 4 µ 4 y = eµη eν, y = y = eµ∂ ψr , y = ∂µψr ∂ ψr . (3.108)

Thus, our new effective Lagrangian is identical to (3.102) except with the replacements

i i AB AB Z → 0, ψa → 0, Y → y . (3.109) 11Notice that ψi and ψ4 began on equal footing, but now we have removed ψi entirely while keeping ψ4. It would be possible to keep them on equal footing, but then we would not be able to employ the guage-fixing procedure and we would therefore have an unnecessarily large number of degrees of freedom. Additionally, the dynamics of ψi versus ψ4 are quite different, so there is no conceptual benefit of treating them on the same footing. 12Notice that we are assuming the quasicrystal space has one additional dimension beyond the physical spatial dimensions for the sake of simplicity. 13 µ µ M Recall that eM ≡ ∂Xr /∂σ .

104 3.5 Higher-form symmetries and the coset construction

When a zero-form symmetry is spontaneously broken, there exists a continuum of vacuum

states. The value of the Goldstone field at a given point A0 can be interpreted as defining

the local vacuum near that point. If we move to a different point, B0, then there will be a

new local vacuum that is related to the local vacuum at A0 by a symmetry transformation. The Maurer-Cartan form tells us how the local vacuum changes from point to point. In

particular, suppose Ω1 is the Maurer-Cartan one-form and V1 is a path connecting A0 and 14 B0. In other words, ∂V1 = A0 ∪ (−B0). We can represent the local vacuum at arbitrary point x by an element of the symmetry group g0(x) ∈ G. Then, we have

∫ Ω V 1 −1 Pe 1 = g0 (B0)g0(A0), (3.110) where P indicates path-ordering. Higher-form symmetries admit a straight-forward generalization of the Maurer-Cartan form. In particular, given a spontaneously broken p-form symmetry, let Ap and Bp be p- dimensional surfaces and Vp+1 be a p + 1-dimensional surfaces such that ∂Vp+1 = Ap ∪ (−Bp).

Then, we define the Maurer-Cartan p + 1-form Ωp+1 by

∫ Ωp+1 Vp+1 −1 e = gp (Bp)gp(Ap), (3.111)

i ∫ ϕp −i ∫ ϕp −1 Qp Ap −1 Qp Bp where gp (Ap) = e and gp (Bp) = e , such that path-ordering is implicit for p = 1, are p-dimensional generalizations of Wilson lines. Here, ϕp is the p-form Goldstone

corresponding to the p-form charge Qp. Notice that for p > 0 we have no notion of path- ordering for the l.h.s., which is related to the fact that all p > 0-form symmetries are abelian; however, if p = 1 then the r.h.s. involves Wilson lines, which must be path-ordered. In either

14 The minus sign on B0 indicates a reversal of orientation.

105 Bp

Vp+1

Ap

Figure 3.2: This figure depicts a p+1-dimensional surface Vp+1 (gray) and two p-dimensional surfaces Ap (red) and Bp (blue). The boundary of Vp+1 is equal to the union of Ap and Bp, that is ∂Vp+1 = Ap ∪ (−Bp). case, we have for p > 0, ∫ Ω V p+1 e p+1 = gp(∂Vp+1), (3.112) where the r.h.s. is path-ordered if p = 1. Supposing we have a mixture of various higher-form symmetries of differing p. Let d d d A = {Ap}p=0, B = {Bp}p=0, and V = {Vp+1}p=0 such that ∂Vp = Ap ∪ (−Bp) for all p. Then, Í we define the Maurer-Cartan to be the mixed-form Ω = p Ωp such that

∫ Ω −1 e V = g (B)g(A), (3.113) where we have left the path-ordering for one-dimensional integrals implicit and we define

¹ Õ ¹ Ö Ö Ω ≡ Ωp, g(A) ≡ gp(Ap), g(B) ≡ gp(Bp). (3.114) V p Vp p p

We are primarily interested in non-equilibrium systems, so we must double field content and define our action on the fluid coordinates σ. As usual we let s = 1, 2 indicate on which leg of the SK contours the fields live. Thus letting the zero-form symmetry generators be

106 15 given by (3.25) and letting Qp be the p > 0-form charges, we have

p ! ∫ µ α A Ö iQp Σ ϕs iX (σ)P¯µ iπ (σ)τα ib (σ)TA gs(Σ) = e p e s e s e s , (3.115) p>0

d where we used Σ = {Σp}p=0 and we identified σ = Σ0. If a given Qp is not spontaneously broken then we impose the time-independent gauge symmetry

∫ p iQp Σ κ gs(Σ) → gs(Σ)· e p , (3.116)

κp = κp(σI) I = 1,..., d p κp = 0 where for is a time-independent -form such that 0M2...Mp . With the group elements (3.115) in mind, we may now compute the Maurer-Cartan mixed-forms

µ ¯ α  A p Ωs = iEs Pµ + i∇sµπ τα + Bs TA + Fs Qp, (3.117)

µ µ p p E = E dσM B A = B A dσM F = 1 F dσM0 ∧ · · · ∧ dσMp where, s sM , s sM and s (p+1)! sM0...Mp . The p > 0-form symmetry associated with charge Qp acts by

∫ p iQp Σ αs (σ) gs(Σ) → e p · gs(Σ), (3.118)

p p for any closed p-forms αs . Notice that since αs can have arbitrary time-dependence, the SK future-time boundary conditions permit the s = 1, 2 transformations to be different; p p p p however, all topological properties of α1 and α2 must be the same. Thus, α1 − α2 must be exact. In this way, there is just one copy of each global p-form symmetry. Finally, once the Maurer-Cartan form is computed, we can go through the usual procedure of constructing manifestly-invariant terms built from the coefficients of the Maurer-Cartan form.

15 We choose Qp to be defined on a surface that “links” with Σp.

107 3.5.1 The zero-temperature limit

There are two meanings of the zero-temperature limit. The first is the most straight- forward. We merely formulate an effective action using the in-out formalism; as a result we only have one set of fields (as opposed to the SK doubled field content). The second meaning, by contrast, involves a thermal system at very low temperature such that all temperature fluctuations can be neglected. Essentially, we are interested in a regime in which thermal activity is suppressed, but not truly absent. As a result, we still need the SK contour and doubled field content. To distinguish these two possibilities, we use T = 0 to indicate the true zero-temperature case and T → 0 to indicate the very small temperature case.

T=0: The coset construction for systems at zero temperature is well-studied, so the only novel comments we can make have to do with higher-form symmetries. With the zero-form generators given by (3.25) and the p > 0-form generators Qp, the most general coset element is ! p ∫ p Ö iQ ϕs µ ¯ α( ) g(Σ) = e Σp eix Pµ eiπ x τα, (3.119) p>0 where we identify x = Σ0 and we have supposed that Qp are spontaneously broken. Then, the Maurer-Cartan mixed-form is defined in the usual way and we can extract from it symmetry- covariant building-blocks. Notice that in the zero-temperature limit, there is no possibility of having unbroken Goldstones or symmetries without corresponding Noether currents.

T→0: Now we take the small-temperature limit. To do this, we retain all of the same machinery as the finite-temperature case, but with two small alterations. We begin by mod- ifying the fluid symmetry (3.28). To determine how to modify this diffeomorphism symmetry, we must first understand where it comes from. Suppose that we have full diffeomorphism symmetry σM → ΣM (σ). For unbroken translations, this is the most natural symmetry to

M µ M suppose because it allows us to gauge-fix by σ = x δµ , which means our EFT is defined

108 on the physical spacetime [35]. At finite temperature, we introduce the inverse temperature four-vector field βM (σ), which transforms as a contravariant vector under diffeomorphisms of the fluid worldvolume. Under a coordinate transformation, we have

∂ΣM βM → βN . (3.120) ∂σN

M M Thus, if we gauge-fix β = β0δ0 , we find that the residual gauge symmetries are (3.28), as expected. But now we wish to take the zero-temperature limit, which is equivalent to taking

M β0 → ∞. As a result, the magnitude of β is no longer meaningful, but the direction is

M M still important. Therefore we impose the gauge-fixing condition β ∝ δ0 , where we allow transformations that alter the magnitude of βM without changing its direction. We are thus left with the residual gauge symmetries

σ0 → σ0 + f (σ0, σI), σI → ΣI(σJ). (3.121)

We take this expanded set of symmetries to be the low-temperature fluid worldvolume sym- metries. The only difference between the T → 0 symmetries and the ordinary fluid symme- tries is that now f may depend arbitrarily on time.

Now we must modify the dynamical KMS symmetries. In the β0 → 0 limit, the trans- formations (3.32) are no longer well-defined. Instead, we impose continuous symmetries of

the form (3.32) or in the classical limit (3.36) but now allow β0 to be a continuous real

parameter. Notice that if we take β0 = θ = 0, we find that our EFT is invariant under Θ alone. Therefore our action is invariant under the following symmetries that do not involve time reversal

0 g1(σ) → g1(σ − iθ, σì ), (3.122) 0 g2(σ) → g2(σ + i(β0 − θ), σì ),

109 for arbitrary θ and β0 such that |θ| ≤ β0. In the classical limit, we have

∂gr gr → gr, ga → ga + iβ0 , (3.123) ∂σ0

for arbitrary β0. Notice that this classical KMS symmetry is a U(1) symmetry that acts only on the advanced fields. As a result, there must be a retarded Noether current that is conserved.16 This Noether current has the interpretation of entropy and is conserved in the

T → 0 limit, as expected. In the T , 0 limit, it is no longer conserved as the dynamical KMS symmetries are no longer continuous [2].

3.5.2 Gauge theories

In this subsection, we will construct the actions for systems involving gauge bosons. We begin by woking at zero-temperature and construct the familiar pure Yang-Mills SU(N) gauge theory. Then we construct the Chern-Simons action, a purely topological theory, from symmetry considerations alone. Finally we turn our attention to finite temperature and reproduce the effective action for magnetohydrodynamics that was first presented in [52].

Yang-Mills

As a simple example, we construct the action for pure Yang-Mills in 3 + 1 dimensions using this new, higher-form coset construction. We consider the zero-temperature, T = 0, case, meaning there is only one copy of the fields. To construct an action with a higher-form symmetry, we must define the Wilson loops. Let ta be the generators of a compact Lie group G such that c [ta, tb] = i fab tc, (3.124)

16The fact that the Noether current is retarded means that it is physical.

110 c where fab are the structure constants. We define the Wilson loop by

iq ∫ A g(C) = Pe C , (3.125)

a µ a where C is a closed path, q is the charge of the Wilson loop, and A ≡ Aµdx t . For Yang- 17 Mills theory, we take G = SU(N); then we will have a global Zn one-form symmetry [69].

Letting S be a two-dimensional surface with ∂S = C, define the Maurer-Cartan two-form Ω2 implicitly by i ∫ Ω iq ∫ A e S 2 = Pe ∂S . (3.126)

Taking S infinitesimal, we can expand both sides of the above equation to find

¹ ¹ 2 ¹ 1 ¹ 1 µ ν q dz dz a b  a b b a  Ω2 = iq A − dλl dλ2 Aµ Aν t t θ(λ1 − λ2) + t t θ(λ2 − λ1) . (3.127) S ∂S 2 0 0 dλ1 dλ2

Using Stoke’s theorem, the r.h.s. can be simplified to yield

¹ ¹ Ω2 = iq (dA + iqA ∧ A), (3.128) S S indicating that Ω2 = iq(dA + iqA ∧ A). Recognize F ≡ dA + iqA ∧ A as the field strength tensor for Yang-Mills theory. At leading order in the derivative expansion, there is only one term that is invariant under the global higher-form symmetry, namely

1 ¹ S = − d4xFa Faµν, (3.129) 4 µν where the factor of −1/4 is a convention that fixes field normalization. We have thus con- structed the familiar Yang-Mills action. 17Ordinarily, the coset construction cannot be used when the symmetry group is discrete as there are no Goldstone modes. There is no guarantee that naïvely applying the coset construction to this case should work, but we will nevertheless proceed and find that everything works out. We should note, however, that unlike the abelian case, the gauge field A does not have the interpretation as a Goldstone as no continuous symmetry is spontaneously broken. These remarks apply equally to the next example in which we consider Chern Simons theory.

111 Chern Simons

One of the original motivations for considering higher-form symmetries was to understand topological phases of matter from symmetry principles alone. Here we will construct the leading-order action for Chern-Simons theory in 2 + 1 dimensions. We will consider a zero- temperature system so we need only one copy of the fields. In addition to Poincaré symmetry, which is unbroken, we postulate the existence of a spontaneously broken one-form symmetry with Goldstone A. Let (3.124) be the generators

a µ a of the compact Lie group G such that A = Aµdx t . The Maurer-Cartan two-form, Ω2, is then defined by18 ∫ ∫ Ω2 A e Σ2 = e ∂Σ2 . (3.130)

Following the Yang-Mills construction, we arrive at the covariant building-block F = dA + A ∧ A, which we interpret as the field strength. Chern-Simons theory involves a term that is symmetry-invariant up to total derivative terms. Unfortunately, the coset construction only furnishes terms that are exactly symmetry- invariant. To circumvent this difficulty, we can construct the necessary term by working in 3 + 1 dimensions and only consider terms that are total derivatives. The leading-order such 3 + 1-dimensional action is then

k ¹ k ¹   2  S3+1 = tr(F ∧ F) = tr d dA + A ∧ A ∧ A , (3.131) 4π M3+1 4π M3+1 3

where M3+1 is a 3 + 1-dimensional manifold and k is a phenomenological constant. Then, using Stoke’s theorem, we obtain the 2 + 1-dimensional action

k ¹  2  S2+1 = tr A ∧ dA + A ∧ A ∧ A , (3.132) 4π M2+1 3 18We drop the factor of i and the charge from the exponent on the r.h.s. so that our conventions match standard results.

112 for 2 + 1-dimensional manifold M2+1. We thus have derived a purely topological field theory from symmetry principles alone!

Magnetohydrodynamics

Magnetohydrodynamics (MHD) is the study of electromagnetism and fluids. In particu- lar, it is the study of fluids that can support the flow of electric current but cannot support electric charge density. Any regions of non-vanishing charge density locally equilibrate to zero exponentially fast. We will restrict our considerations to MHD in 3 + 1 dimensions. Let Aµ

µ be the photon field, Fµν be the field strength given by F = dA, and j be the electromagnetic four-current. Then the electromagnetic equations of motion are

µν µ ∂ν F = j . (3.133)

One might be tempted to say that the symmetries of electromagnetism are the gauged U(1) symmetry, but gauge symmetries are mere redundancies of description and are therefore unphysical. Because this zero-form U(1) symmetry is not a genuine symmetry of nature, it need not survive in our EFT. The true symmetry of electromagnetism is the one-form U(1) symmetry with corresponding current J = ?F, or in index notation,

µν 1 µνλρ J =  Fλρ. (3.134) 2

µν Notice that ∂µJ = 0 is a mathematical identity meaning this current is automatically conserved off-shell. Thus, at the level of the coset construction, we have a one-form symmetry generated by some Q. The SSB pattern is simply that boosts are spontaneously broken and all other symmetries including Q are unbroken. We are only interested in constructing the leading-order action, which, it turns out, factorizes as the difference of two ordinary actions. As a result, we will use just one copy of

113 the fields. The most general group element is

∫ iXµ(σ)P iηi(σ)K iθi(σ)J iQ ϕ g(Σ) = e µ e i e i e Σ2 . (3.135)

Because translations are unbroken, we have the fluid symmetries (3.28) and because rotations and the one-form symmetry are unbroken, we have the gauge symmetries

∫ iλi(σI )J iQ κ g(Σ) → (Σ)· e i, g(Σ) → g(Σ)· e Σ2 , (3.136)

i I I M where λ (σ ) is arbitrary and κ = κM (σ )dσ is a time-independent one-form such that

κ0 = 0. And finally, we have the one-form symmetry ϕ → ϕ + α(σ), for any closed one- form α. From this group element, we find that the Maurer-Cartan form is

µ i i Ω = iE (Pµ + ∇µη Ki) + iΩ Ji + iFQ, (3.137) where

µ ν µ EM = ∂M X [ΛR]ν ,

i −1 M −1 0j ji ∇µη = (E )µ [Λ ∂M Λ] R , 1 (3.138) Ωi = ijk[(ΛR)−1∂ (ΛR)]jk, M 2 M

FMN = ∂[M ϕN],

i i µ iη Ki µ i j iθ Ji ij such that Λ ν = (e ) ν and R = (e ) . We can now impose inverse Higgs constraints to remove the Lorentz Goldstones. First, we remove the boost Goldstones by setting

i E0 = 0, (3.139)

114 which can be solved to give ηi ∂ Xi − 0 tanh η = t , (3.140) η ∂0X p where η ≡ ηì2. Second, to remove the rotation Goldstones, we may fix

ijk −1 M j  (E )i ∂0EM = 0. (3.141)

This inverse Higgs constraint can be used to remove θi from the invariant building-blocks, but the solution is not necessary for the construction of the leading-order action. For the sake of brevity, we therefore will not solve it; interested readers can consult [29] for more details. With these inverse Higgs constraints imposed, the leading-order building-blocks are as follows. Define the fluid worldvolume metric by

µ ν µ ν ∂X ∂X GMN = E ηµν E = ηµν . (3.142) M N ∂σM ∂σN

Then, we have that 1 1 T = t = √ (3.143) E0 −G00

is an invariant building-block, which can be interpreted as the local temperature. Addition- ally, we have the invariant building-block

p MN µ = F0M G F0N, (3.144) where GMN is the inverse of the fluid worldvolume metric. We may interpret µ as the local chemical potential. There are no other invariant building-blocks at this order in the derivative expansion. Performing a coordinate transformation from σM to the physical

115 spacetime x µ ≡ X µ, we have the leading-order action

¹ S = d4xP(T, µ), (3.145) which exactly matches the results of [52]. Finally, we take the T → 0 limit, which can be accomplished by expanding the fluid worldvolume symmetry and allowing full time diffeomorphisms, namely σ0 → σ0 + f (σ0, σI). The only effect this enlarged gauge symmetry has is the removal of T as an invariant building- block. We therefore have ¹ 4 ST→0 = d xP(µ), (3.146) which again matches the results of [52].

3.6 Higher-form symmetries and ’t Hooft anomalies

3.6.1 Gauge fields and the Maurer-Cartan form

We previously stated that the way to ensure all symmetries have corresponding conserved Noether currents is to construct the action using two distinct Maurer-Cartan one-forms—one for each leg of the SK contour—that transform under a single copy of the global symmetry group G. In this section, we will use a Stückelberg trick inspired by [2, 29] to demonstrate that this approach is correct. We begin by introducing sources for the Noether currents corresponding to each symmetry generator of G, which amounts to introducing external gauge fields. Let the zero-form symmetry generators be

P¯m = unbroken translations,

TA = other unbroken generators, (3.147)

τα = broken generators,

116 and let

Qp = p > 0-form charges. (3.148)

We now use m, n = 0,..., d to denote Lorentz indices and µ, ν = 0,..., d to denote physical spacetime coordinate indices.19 We wish to introduce external sources corresponding to each of these symmetries so that we can construct a generating functional for conserved quantities. In particular these external sources are gauge fields corresponding to each symmetry. They are as follows:

m m µ • Let εs (x) = εsµ(x)dx be the vierbeins, which can be thought of as the gauge fields

corresponding to unbroken translations P¯m. The metric tensors are then given by

m n gsµν(x) = εsµ(x)ηmnεsν(x).

A µ • Let As(x) ≡ Asµ(x)dx TA be the gauge fields (or spin connections if the Lorentz group is involved) corresponding to the unbroken zero-form symmetries other than translations.

α µ • Let cs(x) ≡ csµ(x)dx τα be the gauge fields (or spin connections) corresponding to the broken zero-form symmetries.

p p 1 µ0 ∧ · · · ∧ µp • Let Hs = (p+1)! Hsµ0...µp dx dx Qp be the gauge fields corresponding to the p > 0-form symmetries.

On each leg of the SK contour, we can combine these fields into a single object,

m ¯ α A p θs(x) = iεs (x)Pm + ics (x)τα + iAs (x)TA + iHs (x)Qp, (3.149)

0 where the factors of i are included as a matter of convention. Now, letting U(t, t ; θs) for 0 s = 1, 2 be the time-evolution operator from t to t in the presence of external sources θs, the 19It is now necessary to distinguish between Lorentz indices m, n and physical spacetime coordinate indices µ, ν because the Stückelberg trick requires that we gauge all symmetries including Lorentz.

117 generating functional for the conserved currents is

W[θ1,θ2]  †  e = tr U(+∞, −∞; θ1)ρU (+∞, −∞; θ2) . (3.150)

Since θs couple to conserved currents, W[θ1, θ2] must be invariant under two independent copies of the gauge symmetries [2]; that is, for gauge parameters ζ1(x) and ζ2(x) we have

ζ1 ζ2 W[θ1µ, θ2µ] = W[θ1µ, θ2µ]. (3.151)

We can therefore ‘integrate in’ both the broken and unbroken Goldstone fields using the Stückelberg trick. In particular, define

ζs µ α A p Θs(σ) ≡ θs (σ), ζs ≡ {Xs , πs , bs , ϕs }, (3.152) where σM for M = 0,..., d are the fluid worldvolume coordinates. Now we can implicitly define the non-equilibrium effective action by

¹ W[θ1,θ2] µ α A p iIEFT[Θ1,Θ2] e ≡ D[Xs πs s ϕs ] e . (3.153)

m m A α Notice that if we remove the external source fields by fixing εsµ(x) = δµ and Asµ = csµ = p Hs = 0, we find that Θs are nothing other than the Maurer-Cartan mixed-forms (3.117) m m defined on each leg of the SK contour. Since εsµ(x) = δµ we can identify the Lorentz indices

m, n with the physical spacetime coordinate indices µ, ν, allowing us to use P¯µ to refer to the unbroken translation generators. Moreover because these fields live on the SK contour, their values must match up in the infinite future, meaning that even though there were two copies of the gauge fields and gauge symmetries, there is only one copy of the global symmetry group G.

Since W[θ1, θ2] is the generating functional for conserved quantities, we see that any

118 effective action constructed exclusively with building-blocks furnished by the Maurer-Cartan form can be gauged and will have conserved Noether currents associated with each of its symmetry generators.

3.6.2 Dual theories and mixed ’t Hooft anomalies

We now investigate dualities among higher-form currents. Recall the example of the 3+1- dimensional superfluid from §3.2.1 in which there were two conserved charges. One was the U(1) symmetry associated with particle number conservation and the other was associated

µνλ µνλρ µνλ with the higher-form current K =  ∂ρψ. Notice that K is conserved automatically as an identity. However, suppose we gauge the U(1) symmetry by introducing gauge field

Aµ. This can be accomplished by replacing ∂µψ → Aµ + ∂µψ. Then, the higher-form current

µνλ µνλρ becomes K =  (Aρ + ∂ρψ), which is no longer conserved; in particular, we have

µνλ νλ ∂µK = ?F , (3.154) where F = dA. Because the introduction of a gauge field for the U(1) symmetry interferes with the conservation of the higher-form charge, we say that there exists a mixed ’t Hooft anomaly. In particular, it is impossible to gauge the U(1) symmetry and the two-form symmetry simultaneously. But recall how the arguments of the previous subsection relied on the assumption that all symmetries could be gauged simultaneously. Thus, if there is a mixed anomaly, we may only include Goldstones for one of the anomalous symmetries. For the example of the superfluid, we may choose to parameterize the symmetry group element with either the U(1) symmetry generator N and its corresponding Goldstone or the two-form generator Q and its corresponding Goldstone, but not both. It turns out that the presence of a mixed ’t Hooft anomaly is equivalent to SSB. Moreover, this mixed anomaly can be used to prove a kind of Goldstone theorem [68]. Statements regarding mixed anomalies can be generalized to more complicated scenarios.

119 Working in d +1 spacetime dimensions, suppose that Q and R are p-form and d − p−1-form U(1) symmetry generators, respectively and let them have corresponding conserved currents

J µ0...µp and K µ1...µd−p . Suppose further that Q and R enjoy a mixed anomaly such that

if we introduce an external gauge field Hµ0...µp for Q, we have

1 µ1...µd−p µ2...µd−p ν0...νp ∂µ K =  (dH)ν ...ν . (3.155) 1 (p + 1)! 0 p

Then, we may either include the Goldstone ϕ for Q or the Goldstone χ for R in the Maurer- Cartan form, but not both. As we will see, the resulting effective actions will be related by a Legendre transformation in much the same way that the ordinary and dual superfluid actions are related to one another. Supposing that we construct our action with ϕ as the Goldstone field associated with Q. Then, we have that the d − p − 1-form current, which is identically conserved, is given by

K = ?dϕ, (3.156) and the p-form current J , which is conserved on-shell, is obtained by a Noether procedure. Conversely, if we construct an action with χ as the Goldstone associated with R, we find that J = ?dχ (3.157) is identically conserved and K , which is conserved on-shell, is obtained via a Noether proce- dure. Finally, if we are constructing a non-equilibrium EFT, then we must of course double the fields: ϕs and χs for s = 1, 2. To see explicitly how this duality works, suppose we have the non-equilibrium action

IEFT[dϕ1, dϕ2], which depends on the Goldstones ϕs associated with Q and contains no Goldstones associated with R. This action may depend on other fields, but they will not be important for present considerations. Then, we may replace dϕs → Vs for generic p+1-forms

120 Vs and define an auxiliary action by

¹ ¹ IAUX = IEFT[V1, V2] − V1 ∧ dχ1 + V2 ∧ dχ2. (3.158)

Notice that the equations of motion for χs give dVs = 0, which can be solved—assuming a topologically trivial spacetime—to give Vs = dϕs for some p-forms ϕs. We thus recover the original action IEFT. Conversely, we may compute the equations of motion for χs,

δIEFT δIEFT = dχ1, = −dχ2. (3.159) δV1 δV2

20 These equations of motion can then be solved for Vs in terms of dχs and any other fields

that may have appeared in IEFT. We thus arrive at a dual action IDUAL[dχ1, dχ2] with the

Goldstones χs associated with R and no Goldstones associated with Q. Moreover, IEFT and

IDUAL are evidently related by the Legendre transformation (3.158).

3.6.3 Non-equilibrium higher-form ’t Hooft anomalies

To kill the conservation of the higher-form current without killing the higher-form sym- metry, it is important to understand what constitutes a p-form symmetry. Recall that all p-form Goldstones enjoy the time-dependent symmetry (3.118), or equivalently

p p p ϕs → ϕs + αs , (3.160)

p where αs for s = 1, 2 are closed p-forms. These transformations contain within them the global p-form symmetry transformations, but they also contain gauge redundancies. In p p particular, two different sets of transformation parameters αs and α˜s correspond to the p p same physical transformation if and only if their differences αs − α˜s are exact for s = 1, 2. p p Since in the distant future, SK boundary conditions impose ϕ1(+∞) = ϕ2(+∞), the s = 1 20At leading order in the derivative expansion, they can be solved algebraically.

121 and s = 2 transformation parameters must contain the same topological features, meaning p p that α1 − α1 must be exact. Therefore, there is only one copy of the global p-form symmetry and it only acts on the retarded p-form Goldstones. We find it convenient to work in synchronous gauge

ϕp (σ) = 0. s0M2...Mp (3.161)

p p p I Then the residual symmetry of (3.118) now has α1 = α2 = α (σ ) for I = 1,... d, which is now both closed and time-independent. If the higher-form symmetry is unbroken, then we also have the gauge symmetry (3.116). Notice that in either case, the set of symmetries acting on these higher-form fields is time-independent. As a result, the SK boundary conditions— the requirement that the fields be continuous on the closed-time-path21—force us to use the same α for s = 1, 2. Thus, the advanced higher-form fields do no transform under any shift-symmetries, so we expect that there exist advanced building-blocks that contain no derivatives and are invariant under all physical symmetries. It turns out that this is so and we can compute them in the following way. Define the higher-form symmetry generalization of (3.42), denoted by Ga, as ∫ Σp Ga −1 e = g2 (Σp)g1(Σp). (3.162)

Notice that Ga is an element of the Lie algebra as its exponential is a group element. We can therefore express it as a linear combination of symmetry generators

p Ga ⊃ iΦaQp + ··· . (3.163)

p If Qp has a non-equilibrium ’t Hooft anomaly, then we may include Φa as building-block for the action. It may seem strange that we are gauge-fixing the fields before constructing the action.

21 0 If φs are continuous on the closed-time-path, then in the limit that σ → +∞, we have φ1(+∞) = φ2(+∞).

122 Usually, one constructs the action and then gauge-fixes; after all, the components ϕs0M2...Mp yield important constraints as their equations of motion. However, since we are aiming to destroy the conservation of the p-form current, these constraint equations are not desir- able. It is therefore appropriate—indeed necessary—to impose the synchronous gauge-fixing conditions before constructing the effective action.

3.6.4 Quadratic examples

To illustrate some key features of mixed and non-equilibrium ’t Hooft anomalies as applied to systems with higher-form symmetries, we will investigate some simple examples. For concreteness, we will work in 3 + 1 dimensions and suppose the existence of two one-form charges with mixed anomaly, Qel and Qmag, which we interpret as the electric and magnetic one-form charges from electromagnetism. Working at finite temperature 1/β0, and ignoring

µ µ M the conservation of the energy-momentum tensor, we can fix Xs (σ) = δM σ . Then we can µ µ construct our action on the physical spacetime coordinates x ≡ Xs . Let Asµ and ϕsµ be the

Goldstone fields associated with Qel and Qmag, respectively. Since Qel and Qmag enjoy a mixed anomaly, we can never have both As and ϕs appearing in the same effective action. In the A-picture, let

i 0i i 1 i jk jk Fs = dAs E = F , B =  F , (3.164) s s s 2 s

and in the ϕ-picture, let

0i 0i 0i 1 ijk jk fs = dϕs, B = f , E =  f . (3.165) s s s 2 s

We will work in the retarded-advanced basis, but for economy of expression we will drop all r-subscripts on retarded fields. All actions we construct will involve only the relevant and marginal terms and will respect the dynamical KMS symmetry.

123 Hot Maxwell theory

Begin by supposing that Qel and Qmag are both conserved and enjoy a mixed ’t Hooft anomaly; as a result, they are both spontaneously broken. The fact that they are both conserved means that there are no electric or magnetic monopoles. The resulting theory will be the source-free Maxwell theory at finite temperature. Then, we can choose to work in the ϕ- or the A-picture. We choose to work in the A-picture as this choice will highlight the similarities with ordinary electromagnetic theory in vacuum. The action consisting of relevant and marginal terms is

¹ 4  ì ì 2 ì ì I = d x Ea · E − cs Ba · B . (3.166)

The equations of motion are then

ì ì ìÛ 2 ì ì ∇ · E = 0, E = cs ∇ × B. (3.167)

We additionally have the mathematical identities

Û ∇ì · Bì = 0, Bì = −∇ì × Eì. (3.168)

Evidently, there are no electric or magnetic monopoles in this theory. Combining the above equations, we find that ìÜ 2 ì 2 ì ìÜ 2 ì 2 ì B = cs ∇ B, E = cs ∇ E. (3.169)

We thus have a wave that propagates at speed cs. Notice that our equations look just like the vacuum Maxwell equations except the speed of wave-propagation is no longer the standard speed of light. This discrepancy arises because in our example, Lorenz boosts are explicitly broken. Now let us consider the conserved currents associated with the one-form symmetries. We

124 µν µν have Jel and Jmag, where

0i i ij 2 ijk k 0i i ij − ijk k Jel = E , Jel = cs  B , Jmag = B , Jmag =  E . (3.170)

µν µν Then, we have the conservation equations: ∂µJel = 0, which holds on-shell, and ∂µJmag = 0, which holds as an identity. Notice that expressions for the two different currents involve the same building-blocks, allowing us to express the electric higher-form current entirely in terms of components of the magnetic higher-form current, namely

2 cs jk i j J0i = ijk J , J = ijk J0k . (3.171) el 2 mag el mag

We can therefore express the conservation of the electric higher-form current in terms of the components of the magnetic higher-form current by

 2  i jk 0k cs jk 0i  ∂j J + ∂t J = 0, ∂i J = 0. (3.172) mag 2 mag mag

Notice that if Qel and Qmag were unrelated charges, then we would have two independent conserved current, that is there would be no special relationship between the components of the two currents. In this case, we would have two diffusive modes. But in our case, there is a mixed anomaly, which relates the components of the conserved charges. This special rela- tionship (3.171) along with the conservation equations is sufficient to derive the existence of a propagating wave. We therefore see the existence of wave solutions is intimately connected with the existence of a mixed anomaly. Moreover these findings readily generalize to arbitrary dimension and arbitrary p- and d − p−1-form U(1) symmetries with mixed ’t Hooft anomaly.

125 Magnetohydrodynamics I

We will now remove Qel as a physical conserved quantity of the theory by supposing

that Qmag is not spontaneously broken. Physically, the resulting action will correspond to a theory of MHD in which we ignore the fluid degrees of freedom arising from the conservation of the stress-energy tensor. We work in the ϕ-picture and impose the time-independent shift symmetries

ϕsi → ϕsi + κi(xì), (3.173)

for arbitrary κi(xì), which ensure Qmag is unbroken and that Qel is not gauge-invariant (see

below for more details). Thus, Qel is entirely removed from the theory. The action consisting of relevant and marginal terms is

¹   4 D ì0 2 ì0 ì0 ì0 ìÛ0 I = d x i Ea + Ba · B − DEa · E , (3.174) β0 and the resulting equations of motion are

Û ∇ì · Bì0 = 0, Bì0 = D∇ì 2Bì0. (3.175)

We therefore find a diffusion equation for the magnetic field. Compare this result with that of the previous example: when Qmag is spontaneously broken, there exists a propagating wave; when it is unbroken, there is a diffusive mode. This relationship between SSB pattern and dispersion relation in non-equilibrium systems is quite common [29]. µν µν Now consider the conserved currents jel and jmag associated with Qel and Qmag, respec- tively.22 Explicitly, these currents are

0i 0i i j ijk 0k 0i 0i ij ijk Û0k jel = E , jel =  B , jmag = B , jmag = D E . (3.176) 22We used calpital letter J when working in the A-picture and lower-case j when working in the ϕ-picture.

126 µν µν Notice that ∂µ jmag = 0 on the equations of motion and ∂µ jel = 0 identically. It therefore µν seems like there are two conserved one-form symmetries; however jel is not invariant under the gauge transformation (3.173) and is hence not physical. We therefore only have one conserved current, namely jmag.

Magnetohydrodynamics II

We will now consider a gentler way of removing Qel from the set of conserved charges. We will see that we can make the non-conservation of this charge arbitrarily weak by adjusting a time-scale τ. To do this, work in the A-picture and let Qel be both spontaneously broken and have a non-equilibrium ’t Hooft anomaly. As a result, Qmag will still exist, but it will no longer enjoy a mixed anomaly with another conserved charge, which indicates that it is not spontaneously broken. Thus, if we only consider conserved charges, we have the same SSB pattern as in the previous example. The resulting theory will therefore describe MHD, but unlike the more standard formalism, this new theory allows for large charge density.

In order to give Qel a non-equilibrium anomaly, we must operate in synchronous gauge, that is we fix As0 = 0, which yields the residual symmetry

As → As + α(xì), (3.177)

ì for closed, time-independent one-form α such that α0 = 0. Notice that Aa is now a covariant building-block. The action consisting of relevant and marginal terms is therefore

¹   4 ì ì 2 ì ì 1 ì ì i ì2 I = d x Ea · E − cs Ba · B + Aa · E + Aa . (3.178) τ τβ0

The terms with a factor of 1/τ are a result of the non-equilibrium ’t Hooft anomaly. By choosing weak anomaly, we can make this relaxation time τ arbitrarily long.

127 The resulting equations of motion are

Û 1 Eì − c2∇ì × Bì = − Eì. (3.179) s τ

ì Since we have no As0-field in the action, the electric Gauss law is not an equation of motion. Taking the curl and divergence, respectively, of the above equation yields

  ìÜ 2 ì 2 ì 1 ìÛ 1 ì ì B − c ∇ B = − B, ∂t + ∇ · E = 0, (3.180) s τ τ

where we have made use of the identities (3.168). Using the Maxwell relation ∇·ì Eì = ρ, where ρ is the electric charge density, we see that the second equation can be solved by ρ(t, xì) =

−t/τ ρ0(xì)e , meaning that non-zero charge density can exist, but it decays exponentially fast. Let us now investigate the connection to MHD and free Maxwell theory. To recover the

MHD equations of motion, work in the low-frequency limit ∂t  1/τ. Then, the equations of motion become ìÛ 2 ì 2 ì B = τcs ∇ B, ρ = 0. (3.181)

2 If we identify D = τcs , then we recover the diffusive MHD equation (3.175). Further, notice that ρ = 0, indicating that there can be no electric charge density. These equations of motion along with the identities (3.168) are exactly what we expect for standard MHD. To see the

connection with free Maxwell theory, work in the high-frequency limit ∂t  1/τ. Then, we have ìÜ 2 ì 2 ì B = cs ∇ B, ∂t ρ = 0. (3.182)

The first equation indicates that there now exists a propagating wave and the second equation indicates that we are operating on sufficiently short time scales that charge density does not 23 have time to decay. If we take the initial condition ρt=0 = 0, then the the high-frequency 23If we included the conservation of the stress-energy tensor, the second equation of (3.182) would tell us that the charge in each fluid element is ‘locked in’ at leading order in the derivative expansion. If higher-order corrections are considered, this equation becomes diffusive.

128 equations of motion in conjunction with the identities (3.168) reproduce the free Maxwell

equations except the speed of light is replaced with cs.

We arrive at a fascinating result: on short time-scales, Qmag appears unbroken, while on long time-scales, it appears spontaneously broken. Moreover, the characteristic time scale, τ at which this crossover occurs is fixed by the strength of the non-equilibrium ’t Hooft anomaly. We expect this scale-dependent symmetry-breaking to be a much more general phenomenon. Finally, the higher-form currents are given by (3.170), but now only one of them is

conserved. In particular, Jmag is conserved as an identity, while

µi 1 0i ∂µJ = − J . (3.183) el τ el

We see that on time-scales much longer than τ, the r.h.s. of the above equation dominates and we only have one conserved current, just as in the standard MHD example. By con- trast, on time-scales much shorter than τ, the l.h.s. of the above equation dominates and Jel appears essentially conserved, which resembles the hot Maxwell example. Notice that there still exists a relationship among the components of the two higher-form currents (3.171) even though Qel is no longer conserved. We therefore see that in some sense the mixed ’t Hooft anomaly still exists. We thus claim that the correct way to construct a theory of MHD with non-vanishing charge density is to require that Qel and Qmag exhibit a mixed ’t Hooft anomaly and that Qel exhibit a non-equilibrium ’t Hooft anomaly.

Gapped diffusion

We now consider the situation in which both Qel and Qmag are not conserved. Working in the ϕ-picture, we take Qmag to be both unbroken and possess a non-equilibrium ’t Hooft anomaly. Therefore we must impose synchronous gauge (ϕs0 = 0) and endow ϕs with the

129 time-independent gauge symmetries

ϕsi → ϕsi + κi(xì), (3.184)

for arbitrary κi(xì). Notice that ϕai is an allowed building-block. The action consisting of relevant and marginal terms is then

¹   4 D ì0 2 ì0 ì0 ì0 ìÛ0 1 ì0 i 2 I = d x i Ea + Ba · B − DEa · E + ϕìa · B + ϕìa . (3.185) β0 τ τβ0

The equations of motion are therefore

  1 ì0 ì 2 ì0 ∂t + B = D∇ B , (3.186) τ which is a gapped diffusion equation. Thus, there are no gapless excitations.

The higher-form currents are given by (3.176), such that jel is not gauge invariant and is therefore not a physical current. But now, jmag is not conserved, namely

µi 1 0i ∂µ j = − j . (3.187) mag τ mag

Thus, we see explicitly that there are no conserved currents in this theory, which is why there are no gapless excitation.

Legendre transform for unbroken symmetries

We have seen that when a p- and d − p − 1-form current are conserved and enjoy a mixed anomaly, it is possible to perform a Legendre transform to convert the theory involving the p-form Goldstone into a dual theory involving the d − p − 1-form Goldstone. Because these actions are related by a Legendre transform, we can confidently say that they are physically equivalent. By contrast, we merely provided general arguments that the two distinct actions for MHD given by (3.174) and (3.178) were in some sense physically equivalent. Here we

130 will show that there is a duality relating these two actions. ì0 ì0 ì ì Begin by considering (3.174) and replace Es → eìs and Bs → bs, where eìs and bs are fundamental fields. We will work in the retarded-advanced basis and drop the r-subscript on the retarded fields. Notice that eì always appears with a time-derivative, meaning that there is a gauge symmetry eì → eì + ξì(xì), (3.188)

for arbitrary time-independent ξì. Following (3.158), we would like to construct the auxiliary action

¹   ? 4 D 2 ì ì Û ì ì ì ì ì ì IAUX = d x i eìa + ba · b − Deìa · eì + eì · Ea + eìa · E − b · Ba − ba · B , (3.189) β0

ì ì where Es and Bs are given by (3.164). Unfortunately this action contains the term

ì  ì ì  eì · Ea = eì · ∇Aa0 − ∂t Aa , (3.190)

which is not gauge invariant. If, however, we impose Aa0 = 0, then integration by parts removes the gauge non-invariant term. We are then left with the auxiliary action

¹   4 D 2 ì ì Û Û ì ì ì ì ì ì IAUX = d x i eìa + ba · b − Deìa · eì + eì · Aa + eìa · E − b · Ba − ba · B . (3.191) β0

ì Notice that the equations of motion for As and A0 yield

ì ì Û ì ∇ × bs = −eìs, ∇ · eìa = 0, (3.192)

which, with appropriate gauge choice for (3.188), imply that there exist one-forms ϕs such that ì0 ì ì0 eìs = Es, bs = Bs, (3.193)

ì0 ì0 where Es and Bs are given by (3.165). We thus recover the original action (3.174).

131 ì Conversely, to find the dual action, we may instead integrate out eìs and bs. The equations ì for bs are very straight-forward, ì ì bs = Bs. (3.194)

The equations for eìs are

2Di Û ì  ì  eìa − Deì + E = 0, ∂t Deìa − Aa = 0. (3.195) β0

ì ì ì The second equation can be integrated over time to give Deìa + λ(xì) = Aa, where λ is an arbitrary integration function. But the SK boundary condition mandates that all advanced fields vanish in the infinite future, forcing λì = 0, which yields

1 ì eìa = Aa. (3.196) D

And finally, the first equation of (3.195) can be solved to give

Û 2i ì 1 ì eì = Aa + E. (3.197) Dβ0 D

Plugging the solutions (3.194), (3.196), and (3.197) into the auxiliary action (3.191), we obtain the dual action

¹   4 2 ì ì 1 ì ì i ì2 IDUAL = d x −cs Ba · B + Aa · E + Aa , (3.198) τ τβ0

2 ì ì where we have identified D = τcs and re-scaled the fields by As → cs As. At this stage we may gauge-fix A0 = 0 with no consequence as the resulting equations of motion from this component will be entirely redundant. Notice that this dual action exactly matches (3.178) ì ì except for the term Ea · E. The reason this term is missing has to do with power-counting: ì 2 in the original action (3.174), the dynamics are diffusive, so we consider ∂t ∼ ∇ . If τ is ì 2 large, that is 1/τ is of order the energy cutoff for the EFT, then we should expect ∂t ∼ ∇ in

132 ì ì the dual action (3.198). In such a case, the term Ea · E must be considered irrelevant, so we ought not include it in the leading-order action. By contrast, when we constructed (3.178) from symmetry considerations, we assumed that the non-equilibrium ’t Hooft anomaly was weak, meaning that 1/τ is much less than the cutoff energy. As a result, we may consider ì ì ì ∂t ∼ ∇, making the term Ea · E marginal.

3.7 Dual superfluids

In this section, we will construct the dual two-form EFT for 3+1-dimensional superfluids at zero and finite temperature. For a coset construction of the ordinary superfluid action, consult [29].

3.7.1 Zero temperature

In the dual description of the superfluid, we have, in addition to Poincaré symmetry, the U(1) two-form symmetry generator Q and U(1) zero-form generator N, which counts particle number. Moreover, there is a mixed anomaly between Q and N, meaning that we can only include one of them in our coset construction. First, suppose we take N to be in our coset. Then, the standard claim is that both Q

and P0 are spontaneously broken, but their diagonal subgroup P¯0 = P0 + µN is preserved [51, 50]. Boosts are also spontaneously broken but all other symmetries are preserved. Thus the most general coset element takes the form

µ ¯ i g(x) = eix Pµ eiπ(x)N eiη (x)Ki . (3.199)

Notice that we can rewrite this group element as

µ i g(x) = eix Pµ eiψ(x)N eiη (x)Ki, ψ ≡ µt + π. (3.200)

Thus, we can forget about the fact that P0 is broken; instead we may act as though only

133 N is broken and then after the fact give ψ a time-dependent VEV. The coset construction for the ordinary superfluid action is already well-studied, so we will now change gears and consider the dual theory. In the dual picture, we take Lorentz boosts and Q to be spontaneously broken and we suppose our system exists at finite particle-number density. Because we are constructing an EFT describing a zero-temperature system, we need not bother with the SK formalism, so there is just one copy of each field. Further, we define our coset on the physical spacetime directly and we only include Goldstones associated with broken symmetries. The most general coset element is ∫ ixµ P iηi(x)K iQ A g(Σ) = e µ e i e Σ2 , (3.201) where Σ = {x, Σ2}, for two-dimensional manifold Σ2, and where A is a two-form field. Notice that in the dual picture, we do not treat P0 as spontaneously broken. Much like (3.200) we must instead give A(x) a spacetime-dependent VEV. It turns out that requiring

k Aµν = Aµν + ϕµν, hA0ii = 0, Ai j ∝ i jk x , (3.202) is equivalent to having a dual theory at finite particle-number density [78]. Thus, it is the dual-theory equivalent of having P0 and N broken but their diagonal subgroup preserved in the original picture (3.199). The resulting Maurer-Cartan form is

ν i i Ω = iE (Pν + ∇νη ) + ω Ji + ?FQ, (3.203)

134 where

ν ν Eµ = Λ µ,

i −1 ν −1 0i ∇µη = (E )µ[Λ ∂νΛ] , 1 (3.204) ωi = i jk[Λ−1∂ Λ]jk, µ 2 µ 1 F µ =  µνλρ∂ A . 2 ν λρ

µ ν µ It is convenient to also define F = F Eν , which is boost-invariant. Now impose inverse Higgs constraints to remove the boost Goldstones. In particular, fix

0 = F i. (3.205)

(ν) ν The quickest way to solve this equation is to note that eµ = Λµ form a set of orthonormal 0 (ν) µµ0 (ν ) νν0 µ vectors such that eµ η eµ0 = η . Thus, the inverse Higgs constraints imply that F is (i) (0) 0 orthogonal to eµ and hence parallel to eµ = Λµ . We therefore have that

0 Fµ Λµ = , (3.206) F

p µ i where F = −F Fµ. In terms of the boost Goldstones η , we have

ηi Fi tanh η = , (3.207) η F0

p 2 0 0 i where η = ηì and we have used the fact that Λ0 = cosh η and Λi = (η /η) sinh η. Notice

that this equation is only non-singular because we supposed that Aµν enjoys the VEV (3.202). With these inverse Higgs constraints solved, we have just one invariant building-block at leading order in the derivative expansion, namely

t p µ Y ≡ F = −F Fµ. (3.208)

135 The leading-order effective action is therefore

¹ S = d4xL(Y), (3.209) for some function L. Notice that the zero-form charge N associated with particle number conservation did not appear anywhere in this coset construction.24 The reason is that in the dual picture, the conserved current appears as µ 1 µνλρ J =  ∂ν Aλρ, (3.210) U(1) 2 which is conserved off-shell as a mathematical identity.

3.7.2 Finite temperature

We now turn our attention to the finite temperature case. The SSB patter is of course the same. The difference is that now we define the action on the fluid worldvolume and parameterize the full symmetry group with Goldstones even if they are associated with unbroken symmetries. At finite temperature, we define our action on the SK contour and as a result must have doubled field content. If, however, we are only interested in the leading- order action, then it turns out that the dynamical KMS symmetries force the non-equilibrium action to factorize as the difference of two ordinary actions, each with a single copy of the fields [29]. We will therefore work with just one copy of the fields. Higher-order corrections require two copies of the fields. The most general group element is

∫ iXµ(σ)P iηi(σ)K iθi(σ)J iQ A g(Σ) = e µ e i e i e Σ2 , (3.211)

24Similarly in the ordinary coset construction, the two-form symmetry does not appear.

136 where Σ = {σ, Σ2}, for two-dimensional manifold Σ2. Then the Maurer-Cartan form is

µ i i Ω = iE (Pµ + ∇µη Ki) + iω Ji + i ? FQ, (3.212)

where

µ ν µ EM = ∂M X [ΛR]ν ,

i −1 M −1 0j ji ∇µη = (E )µ [Λ ∂M Λ] R , 1 (3.213) ωi = i jk[(ΛR)−1∂ (ΛR)]jk, M 2 M 1 F M =  MNPQ∂ A , 2 N PQ

i i µ iη Ki µ i j iθ Ji ij such that Λ ν = (e ) ν and R = (e ) . We can now impose inverse Higgs constraints to remove the Lorentz Goldstones. First, to remove boost Goldstones, we impose (3.139), which can be solved to yield

ηi ∂ Xi − 0 tanh η = t , (3.214) η ∂0X

p where η ≡ ηì2.25 Second, to remove the rotation Goldstones, fix the constraint (3.141). As mentioned previously, we need not solve this inverse Higgs constraint for leading-order actions. After imposing these inverse Higgs constraints, the leading-order building-blocks are as follows. Begin by defining the fluid worldvolume metric by (3.142). Then, we have the local temperature T given by (3.143) as an invariant building-block. Finally, we have that

p M N M Y = F GMN F , Z = F GM0, (3.215)

are the remaining invariant building-blocks. Thus, transforming to the phsyical spaceitme

25 M −1 i We could have alternatively imposed inverse Higgs constraints F (E )M = 0, in keeping with (3.205). At least at leading-order these two inverse Higgs constraints end up yielding the same EFT.

137 coordinates x µ = X µ, the leading-order action is

¹ S = d4xL(T,Y, Z). (3.216)

3.8 Dual solids

In this section, we will construct the dual two-form EFTs for solids at zero and finite temperature. The two form charges count the number of crystal defects of the lattice.

3.8.1 Zero temperature

At zero temperature, the standard EFT for a solid involves three Goldstone modes φi for i = 1, 2, 3, corresponding to spontaneously broken translations [1, 29, 77, 57]. At leading order in the derivative expansion, we have

¹ 4 ij ij i µ j S = d xP(γ ), γ = ∂µφ ∂ φ , (3.217) for some function P. Note that these Goldstones are invariant under the shift symmetries

φi → φi + ci, (3.218) for constant ci. Thus this effective field theory has three internal U(1) symmetries, which have the interpretation of lattice momentum. Let Pi be the generators of these U(1) symmetries. We additionally have three distinct two-form symmetries with conserved currents

ìµνλ µνλρ J =  ∂ρφ.ì (3.219)

It is the aim of this subsection to take these two-form symmetries as the starting point. Recall that SSB pattern of solids is as follows. Boosts and rotations are spontaneously broken. Physical spatial translations and lattice-momentum translations are spontaneously

138 broken, but their diagonal subgroup Pi + Pi are preserved, and temporal translations are preserved. In the dual picture, we do not make use of the lattice momentum generators Pi

and hence we cannot treat them or Pi as spontaneously broken. We must, however, still continue to treat boosts and rotations as spontaneously broken. In the dual picture, in addition to Poincaré symmetry, we have a three-vector of spon- taneously broken two-form symmetries with generator Qì that enjoy mixed anomalies with Pì. As a result, if we include Goldstones for Qì, we cannot have Goldstones for Pì in the coset. Since we are working at zero temperature, we only need one copy of the fields, we can formulate our EFT directly on the physical spacetime manifold, and we need only include Goldstones for spontaneously broken symmetries. The most general coset element is

ì ∫ ixµ P iηi(x)K iθi(x)J iQ· ϕì g(Σ) = e µ e i e i e Σ2 , (3.220)

where Σ = {x, Σ2}, for two-dimensional manifold Σ2. Then the resulting Maurer-Cartan form is ν i i ì ì Ω = iE (Pν + ∇νη + ∇νθ Ji + ?F · Q), (3.221) where

ν ν Eµ = (ΛR) µ,

i −1 ν −1 0i ∇µη = (E )µ[(ΛR) ∂ν(ΛR)] , 1 (3.222) ∇ θi = ijk(E−1)ν [(ΛR)−1∂ (ΛR)]jk, µ 2 µ ν ì µ µ ìν F = Eν F ,

i i ìµ 1 µνλρ µ iη Ki µ ij iθ Ji ij such that F ≡ 2  ∂ν ϕìλρ, Λ ν = (e ) ν, and R = (e ) . We can now impose inverse Higgs constraints to remove the Lorentz Goldstones. To

139 remove the rotation Goldstones, we impose26

ijk F jk = 0, (3.223)

which can be solved to give

ij 1/2 ij i j iµ j F = (Y ) , Y = F Fµ, (3.224) where Y 1/2 is the matrix square root of Y. Then, to remove the boost Goldstones, we impose

F it = 0, (3.225) which can be solved to give ηi tanh η = −F jt(Y −1/2)ji. (3.226) η

We are thus left with only one set of invariant building-blocks, namely Yij. The leading-order dual action is then ¹ 4 ij SDUAL = d xL(Y ), (3.227) for some function L. Notice how both the standard solid action and the dual solid action depend on symmetric 3 × 3 matrices. Finally, as in the superfluid case, we should expect the two-form fields to have non-trivial background values. In particular, we want Fiµ ∝ ηiν. We therefore suppose that D i E D i E i jk ϕ0j = 0, ϕjk ∝  t. (3.228) 26The indices i, j = 1, 2, 3 are now playing two roles: the first is the role of spatial rotation index and the second is to enumerate the components of Qì. We can get away with this redundant notation because rotations are spontaneously broken.

140 3.8.2 Finite temperature

At finite temperature, the symmetry-breaking pattern is unchanged; however, lattice mo-

mentum generated by Pi need not be conserved. If it is conserved, we have the phenomenon of second sound; otherwise there is no second sound [32, 59]. Since our EFTs now exist at finite temperature, we have two copies of the fields, in

i particular, we have two copies of the solid fields φs for s = 1, 2. Alternatively, we have in i i i the retarded-advanced basis, φr and φa. If lattice momentum is conserved, then both φr and i i i φa must always appear in the action with at least one derivative, namely ∂µφr and ∂µφa. i If, however, lattice momentum is not conserved, but Pi is still a symmetry, then φa may i i appear without derivatives even as φr must always appear with a derivative. Recall that φr i are the classical fields, while φa encode information about fluctuations. Thus, whether or not lattice momentum is conserved, we have the three-form conserved currents

ìµνλρ µνλρ J =  ∂ρφìr . (3.229)

27 If we then switch to the dual picture and work with two-form fields ϕìsµν, the currents associated with Pi are now given by

ìµ 1 µνλρ J =  ∂ν ϕìrλρ. (3.230) 2

ìµ But notice that ∂µJ = 0 identically because partial derivatives commute. It therefore seems that the dual picture automatically has conserved lattice momentum; however, this need not be true if J µ is not gauge invariant. Notice that if Qì are unbroken, then we have the gauge symmetries of the form (3.116), which do not leave J µ invariant. Thus, from the dual perspective, we have conserved lattice momentum and hence second sound if and only if Qì are spontaneously broken.

27 µ We can take ϕìsµν to be the pullback of ϕìsM N to the physical spacetime via the inverse of the map Xs (σ).

141 We now proceed to constructing the dual actions for solids with and without second sound. As in the zero-temperature case, we have spontaneously broken boosts and rotations, but we do not treat translations as broken.28 We also have the three-form symmetry charges Qì. The most general group element is

ì ∫ iXµ(σ)P iηi(σ)K iθi(σ)J iQ· ϕì g(Σ) = e µ e i e i e Σ2 , (3.231)

where Σ = {σ, Σ2}, for two-dimensional manifold Σ2. And the Maurer-Cartan form is

µ i i ì ì Ω = iE (Pµ + ∇µη Ki + ∇µθ Ji) + i ? F · Q, (3.232) where

µ ν µ EM = ∂M X [ΛR]ν ,

i −1 M −1 0j ji ∇µη = (E )µ [Λ ∂M Λ] R , 1 (3.233) ∇ θi = i jk(E−1)M [(ΛR)−1∂ (ΛR)]jk, µ 2 µ M 1 FìM =  MNPQ∂ ϕì , 2 N PQ

i i µ iη Ki µ i j iθ Ji ij iµ µ iM such that Λ ν = (e ) ν and R = (e ) . It will be convenient to define F = EM F . The inverse Higgs constraints necessary to remove boost Goldstones are

ηi Ei = 0 =⇒ tanh η, (3.234) 0 η

and the inverse Higgs constraints necessary to remove rotation Goldstones are

i jk F jk = 0 =⇒ F ij = (y1/2)ij, (3.235)

28Instead of treating spatial translations as spontaneously broken, we endow the two-form fields with the VEVs (3.228)

142 ji iµ j where y = F Fµ. The invariant building-blocks are as follows. First, we have the local temperature given by (3.143). Next, we have the symmetric matrix yi j as defined above; notice that both of these terms are invariant under the gauge symmetry

K ϕIJ(σ) → ϕIJ(σ) + κIJ(σ ). (3.236)

Thus these building-blocks are invariant whether or not Qì are spontaneously broken. Lastly, we have the building-block

∂X µ zi ≡ F it = uµFi, where uµ ≡ T , (3.237) µ ∂σ0

but it is not invariant under the gauge symmetry (3.236). Thus it is only an allowed building- block if Qì are spontaneously broken, that is, when lattice momentum is conserved. Transforming to physical spacetime, the dual actions for solids are therefore

¹ 4 ij i SDUAL = d xL(T, y , z ), (3.238) describing a solid with conserved lattice momentum, and

¹ 4 i j SDUAL = d xL(T, y ), (3.239) describing a solid without conserved lattice momentum. The building-block zi captures the relative motion of the fluid29 and the lattice when second sound is present. 29Second sound emerges when a thermal cloud of solid phonons can flow as a fluid independently of the solid lattice [59, 79]. Then we have both fluid sound waves and solid (i.e. lattice) sound waves.

143 3.9 Discussion

In this chapter, we presented a systematic procedure to formulate non-equilibrium ef- fective actions for hydrodynamic (gapless) and quasi-hydrodynamic (weakly-gapped) exci- tations involving p-form symmetries. To aid in the systematization, we employed the coset construction, generalizing it to account for higher-form symmetries in non-equilibrium sys- tems. In addition, we extended the coset construction to account for theories exhibiting non-equilibrium ’t Hooft anomalies, which are exact symmetries of the action but have no conserved Noether current. As the name suggests these anomalies can only occur in non- equilibrium systems. Further, we noticed some interesting features of non-equilibrium EFTs that involve higher-form symmetries:

• When implementing the coset construction, the Maurer-Cartan form associated with

a p-form symmetry, which we denote by Ωp+1, is a p + 1-form. If various higher-form symmetries of different p exist in a system, then the resulting Maurer-Cartan form is Í a mixed-form in the sense that it can be expressed as a sum of the form Ω = p Ωp+1.

• Whenever a U(1) p-form symmetry is spontaneously broken, there exists a d−p−1-form U(1) symmetry that is also spontaneously broken. Further, the p-form and d − p − 1- form symmetries exhibit a mixed ’t Hooft anomaly (not to be confused with a non- equilibrium ’t Hooft anomaly). Moreover, if there exist p-form and a d−p−1-form U(1) symmetries with mixed ’t Hooft anomaly, then both higher-form symmetries must be spontaneously broken. Such a phenomenon holds both at zero and finite temperature.

• Whenever such a mixed anomaly exists, it is impossible to have the p-form and the d − p−1-form Goldstones in the same action. There are therefore dual physically equivalent descriptions. In particular, one effective action involves the p-form Goldstone and the other involves the d − p − 1-form Goldstone.

144 • In non-equilibrium systems, Goldstone-like excitations exist even for unbroken p-form symmetries. The distinguishing feature between spontaneously broken and unbroken p-form Goldstones is the emergence of a gauge symmetry. These findings are a natural extension of those in [29]. The conserved d − p − 1-form charge that would exist if the p-form symmetry were spontaneously broken becomes gauge non-invariant and thus fails to be a physical charge.

• There is a second way to construct an action with an unbroken U(1) p-form symmetry. We work in the dual picture involving the d − p − 1-form Goldstone and give it a non- equilibrium ’t Hooft anomaly. This leads to non-conservation of the d − p − 1-form current but does not affect the conservation of the p-form current. As a result, the p-form symmetry, as it enjoys no mixed ’t Hooft anomaly, cannot be spontaneously broken. Further, we can give the d − p − 1-form current an arbitrarily weak non- equilibrium ’t Hooft anomaly, meaning that we can, in a certain sense, have a weakly unbroken p-form symmetry.

• Lastly, using the coset construction, we were able to reproduce the well-known Chern- Simons action. That is, from symmetry principles alone, we were able to construct a purely topological theory.

This work admits generalizations and applications in many directions. First, there is an interesting consequence of these findings that warrants further investigation: in non- equilibrium systems, whether or not a given symmetry is spontaneously broken may depend on the time-scale over which it is observed. Consider the situation of a p- and a d − p − 1- form symmetry with mixed ’t Hooft anomaly. If we give the d − p − 1-form symmetry a non-equilibrium ’t Hooft anomaly, the strength of the anomaly is characterized by a time- scale τ in the sense that for frequency ωτ  1, the current is not conserved, whereas for ωτ  1, the current appears conserved. In this way, the non-equilibrium ’t Hooft anomaly clearly exists in the low-frequency regime, meaning that the p-form symmetry is

145 unbroken. However in the high-frequency regime, the non-equilibrium ’t Hooft anomaly is highly suppressed, meaning that the p-form symmetry will appear broken. In this chapter, we saw that our quadratic models for electromagnetic systems behave in this way, but it appears to be a more general phenomenon. Second, the existing non-equilibrium effective actions for magnetohydrodynamic systems do not account for situations in which there is a build-up of electric charge. By exploiting principles of mixed and non-equilibrium ’t Hooft anomalies, we were able to construct a quadratic magnetohydrodynamic action that permits large electric charge density, which exponentially decays to zero density. Such an effective action may ultimately yield important insights into the nature of instabilities, which are of great importance in fusion reactions. Third, higher-form symmetries are often associated with topological phases of matter. The non-equilibrium coset construction involving higher- form symmetries may therefore prove to be a powerful tool for constructing theories for topological phases beyond the Chern-Simons action. Fourth, we have only considered one kind of mixing between various p-form symmetries, namely we considered mixed ’t Hooft anomalies. However, there are other, more complicated relationships that can exist among higher-form symmetries, like two-groups [76]. Extending the coset construction to account for objects such as two-groups is of significant theoretical interest. Finally, higher-form symmetries have many applications in a wide variety of areas. For example, they can be used to construct effective field theories of superfluid vortices [78] and we expect they should play a similar role in describing the dynamics of disclinations in solids in the non-equilibrium EFT formalism [80].

146 Epilogue

The effective field theory approach to many-body physics provides new lines of attack for understanding a wide range of topics. As this EFT approach is based in symmetry principles, all guess-work is removed, so it is no longer necessary to rely on ad hoc assumptions about local thermal equilibrium in order to constrain the terms of the action. Instead, everything is fixed—up to finitely many phenomenological coefficients at any given order in the derivative expansion—by symmetry considerations. In this thesis, we have vastly expanded the number of systems that can be described by non-equilibrium EFTs, that is EFTs defined on the Schwinger-Keldysh contour that can account for dissipation and noise. Previously the only such known actions existed in some sort of fluid phase, e.g. ordinary fluids, superfluids, and string fluids. Here, we extended the list of EFTs to include those that describe solids, supersolids, nematic liquid crystals, smectic liquid crystals in phases A, B, and C, chemically reacting fluids, quasicrystals, higher- form dual theories of superfluids and solids, and plasmas that can support large charge density. Moreover, by way of extending the coset construction to the realm of non-equilibrium phenomena, we have a very general way to construct EFTs for a wide range of systems. In particular, we have identified the rules for constructing EFTs with both ordinary symmetries and higher-form symmetries. Using the coset construction, we saw that the low-energy modes of systems at finite- temperature are not always the traditional Goldstones associated with spontaneous sym- metry breaking. Instead, we invoked the intuition that while the density matrix as a whole

147 might leave a given symmetry unbroken, the individual microstates of the ensemble described by the thermal density matrix will in general be highly chaotic and hence will spontaneously break every symmetry. We therefore saw fit to include Goldstones associated with every symmetry, even unbroken ones. But to distinguish between traditional Goldstones associ- ated with broken symmetries and these new unbroken Goldstones, we required that unbroken Goldstones enjoy a kind of time-independent gauge symmetry. The inclusion of these addi- tional Goldstones required the invention of new inverse Higgs constrains to remove extraneous Goldstones. Further, we found that certain inverse Higgs constraints could be imposed that can remove second sound from certain EFTs. In particular, these inverse Higgs constraints account for Umklapp scattering, which is the process by which lattice momentum is not conserved in solids. At the level of the EFT, lattice momentum is the conserved quantity associated with the emergent internal translational symmetry. This symmetry is absolutely necessary if the solids are to have any sort of large-scale homogeneity. We found that it is possible to keep the global symmetry, while killing the conservation of the corresponding Noether current; this is evidently not possible for ordinary EFTs. The possibilities for applications and extensions are many. First, the non-equilibrium EFT approach has so far been focused mainly to attacking problems in which the equilib- rium state of the system is thermal. In biophysics, however, there are may examples in which such an assumption is explicitly violated as energy from the environment enters a biological system and performs work. Such systems are known as active matter systems. Extending the formalisms presented in this thesis could yield exciting results for active bio- physical systems. Second, our novel approaches to plasmas hint at exciting possibilities. In particular, unlike previous attempt to construct EFTs for plasmas, our action can be easily generalized to describe many different kinds of systems. A most interesting possibility is the construction of an EFT for quark-gluon plasma. These systems have repelled most attempts at analytic description, but the powerful machinery of the non-equilibrium EFT paradigm

148 might provide useful analytic tools. Finally, the non-equilibrium EFT approach is highly constrained; as mentioned earlier, it removes all need of guess-work. In this way, reformu- lating existing theories with this approach may reveal missing elements and demonstrate connections previously unnoticed.

149 References

[1] A. Nicolis, R. Penco, F. Piazza, and R. Rattazzi, “Zoology of condensed matter: Framids, ordinary stuff, extra-ordinary stuff,” JHEP, vol. 06, p. 155, 2015. arXiv: 1501.03845 [hep-th].

[2] H. Liu and P. Glorioso, “Lectures on non-equilibrium effective field theories and fluctu- ating hydrodynamics,” PoS, vol. TASI2017, p. 008, 2018. arXiv: 1805.09331 [hep-th].

[3] P. Glorioso, M. Crossley, and H. Liu, “Effective field theory of dissipative fluids (II): classical limit, dynamical KMS symmetry and entropy current,” JHEP, vol. 09, p. 096, 2017. arXiv: 1701.07817 [hep-th].

[4] M. Crossley, P. Glorioso, and H. Liu, “Effective field theory of dissipative fluids,” JHEP, vol. 09, p. 095, 2017. arXiv: 1511.03646 [hep-th].

[5] P. Glorioso, H. Liu, and S. Rajagopal, “Global Anomalies, Discrete Symmetries, and Hydrodynamic Effective Actions,” JHEP, vol. 01, p. 043, 2019. arXiv: 1710.03768 [hep-th].

[6] P. Gao, P. Glorioso, and H. Liu, “Ghostbusters: Unitarity and Causality of Non- equilibrium Effective Field Theories,” JHEP, vol. 03, p. 040, 2020. arXiv: 1803.10778 [hep-th].

[7] P. Glorioso and H. Liu, “The second law of thermodynamics from symmetry and uni- tarity,” Dec. 2016. arXiv: 1612.07705 [hep-th].

[8] M. Blake, H. Lee, and H. Liu, “A quantum hydrodynamical description for scrambling and many-body chaos,” JHEP, vol. 10, p. 127, 2018. arXiv: 1801.00010 [hep-th].

[9] S. Dubovsky, L. Hui, A. Nicolis, and D. T. Son, “Effective field theory for hydrodynam- ics: thermodynamics, and the derivative expansion,” Phys. Rev. D, vol. 85, p. 085 029, 2012. arXiv: 1107.0731 [hep-th].

[10] A. Nicolis, R. Penco, F. Piazza, and R. A. Rosen, “More on gapped Goldstones at finite density: More gapped Goldstones,” JHEP, vol. 11, p. 055, 2013. arXiv: 1306.1240 [hep-th].

[11] A. Nicolis, R. Penco, and R. A. Rosen, “Relativistic Fluids, Superfluids, Solids and Supersolids from a Coset Construction,” Phys. Rev. D, vol. 89, no. 4, p. 045 002, 2014. arXiv: 1307.0517 [hep-th].

150 [12] A. Nicolis, “Low-energy effective field theory for finite-temperature relativistic super- fluids,” Aug. 2011. arXiv: 1108.2513 [hep-th].

[13] M. Harder, P. Kovtun, and A. Ritz, “On thermal fluctuations and the generating functional in relativistic hydrodynamics,” JHEP, vol. 07, p. 025, 2015. arXiv: 1502. 03076 [hep-th].

[14] N. Banerjee, J. Bhattacharya, S. Bhattacharyya, S. Jain, S. Minwalla, and T. Sharma, “Constraints on Fluid Dynamics from Equilibrium Partition Functions,” JHEP, vol. 09, p. 046, 2012. arXiv: 1203.3544 [hep-th].

[15] K. Jensen, N. Pinzani-Fokeeva, and A. Yarom, “Dissipative hydrodynamics in super- space,” JHEP, vol. 09, p. 127, 2018. arXiv: 1701.07436 [hep-th].

[16] K. Jensen, R. Marjieh, N. Pinzani-Fokeeva, and A. Yarom, “A panoply of Schwinger- Keldysh transport,” SciPost Phys., vol. 5, no. 5, p. 053, 2018. arXiv: 1804 . 04654 [hep-th].

[17] ——, “An entropy current in superspace,” JHEP, vol. 01, p. 061, 2019. arXiv: 1803. 07070 [hep-th].

[18] K. Jensen, M. Kaminski, P. Kovtun, R. Meyer, A. Ritz, and A. Yarom, “Towards hydrodynamics without an entropy current,” Phys. Rev. Lett., vol. 109, p. 101 601, 2012. arXiv: 1203.3556 [hep-th].

[19] P. Kovtun, G. D. Moore, and P. Romatschke, “Towards an effective action for rela- tivistic dissipative hydrodynamics,” JHEP, vol. 07, p. 123, 2014. arXiv: 1405.3967 [hep-ph].

[20] S. Grozdanov and J. Polonyi, “Viscosity and dissipative hydrodynamics from effec- tive field theory,” Phys. Rev. D, vol. 91, no. 10, p. 105 031, 2015. arXiv: 1305.3670 [hep-th].

[21] F. M. Haehl, R. Loganayagam, and M. Rangamani, “Two roads to hydrodynamic effective actions: a comparison,” Jan. 2017. arXiv: 1701.07896 [hep-th].

[22] ——, “Effective Action for Relativistic Hydrodynamics: Fluctuations, Dissipation, and Entropy Inflow,” JHEP, vol. 10, p. 194, 2018. arXiv: 1803.11155 [hep-th].

[23] ——, “The Fluid Manifesto: Emergent symmetries, hydrodynamics, and black holes,” JHEP, vol. 01, p. 184, 2016. arXiv: 1510.02494 [hep-th].

[24] M. Hongo, S. Kim, T. Noumi, and A. Ota, “Effective Lagrangian for Nambu-Goldstone modes in nonequilibrium open systems,” Jul. 2019. arXiv: 1907.08609 [hep-th].

151 [25] T. Hayata, Y. Hidaka, T. Noumi, and M. Hongo, “Relativistic hydrodynamics from quantum field theory on the basis of the generalized Gibbs ensemble method,” Phys. Rev. D, vol. 92, no. 6, p. 065 008, 2015. arXiv: 1503.04535 [hep-ph].

[26] M. Hongo, “Path-integral formula for local thermal equilibrium,” Annals Phys., vol. 383, pp. 1–32, 2017. arXiv: 1611.07074 [hep-th].

[27] ——, “Nonrelativistic hydrodynamics from quantum field theory: (i) normal fluid com- posed of spinless schrödinger fields,” Jan. 2018. arXiv: 1801.06520 [cond-mat.stat-mech].

[28] D. V. Volkov, “Phenomenological Lagrangians,” Fiz. Elem. Chast. Atom. Yadra, vol. 4, pp. 3–41, 1973.

[29] M. J. Landry, “The coset construction for non-equilibrium systems,” JHEP, vol. 07, p. 200, 2020. arXiv: 1912.12301 [hep-th].

[30] ——, “Dynamical chemistry: non-equilibrium effective actions for reactive fluids,” Jun. 2020. arXiv: 2006.13220 [hep-th].

[31] M. Baggioli and M. Landry, “Effective Field Theory for Quasicrystals and Phasons Dynamics,” SciPost Phys., vol. 9, no. 5, p. 062, 2020. arXiv: 2008.05339 [hep-th].

[32] M. J. Landry, “Second sound and non-equilibrium effective field theory,” Aug. 2020. arXiv: 2008.11725 [hep-th].

[33] ——, “Higher-form symmetries and ’t Hooft anomalies in non-equilibrium systems,” Jan. 2021. arXiv: 2101.02210 [hep-th].

[34] L. Hui, A. Joyce, M. J. Landry, and X. Li, “Vortices and waves in light dark matter,” JCAP, vol. 01, p. 011, 2021. arXiv: 2004.01188 [astro-ph.CO].

[35] M. J. Landry and G. Sun, “The coset construction for particles of arbitrary spin,” Oct. 2020. arXiv: 2010.11191 [hep-th].

[36] M. Baggioli, M. Landry, and A. Zaccone, “Deformations, relaxation and broken sym- metries in liquids, solids and glasses: a unified topological field theory,” Jan. 2021. arXiv: 2101.05015 [cond-mat.soft].

[37] A. Esposito, R. Krichevsky, and A. Nicolis, “Vortex precession in trapped superfluids from effective field theory,” Phys. Rev. A, vol. 96, no. 3, p. 033 615, 2017. arXiv: 1704. 08267 [hep-th].

[38] A. Caputo, A. Esposito, F. Piccinini, A. D. Polosa, and G. Rossi, “Directional detec- tion of light dark matter from three-phonon events in superfluid 4He,” Phys. Rev. D, vol. 103, no. 5, p. 055 017, 2021. arXiv: 2012.01432 [hep-ph].

152 [39] A. Caputo, A. Esposito, and A. D. Polosa, “Light Dark Matter and Superfluid He-4 from EFT,” J. Phys. Conf. Ser., vol. 1468, no. 1, M. Nakahata, Ed., p. 012 060, 2020. arXiv: 1911.07867 [hep-ph].

[40] A. Caputo, A. Esposito, E. Geoffray, A. D. Polosa, and S. Sun, “Dark Matter, Dark Photon and Superfluid He-4 from Effective Field Theory,” Phys. Lett. B, vol. 802, p. 135 258, 2020. arXiv: 1911.04511 [hep-ph].

[41] A. Caputo, A. Esposito, and A. D. Polosa, “Sub-MeV Dark Matter and the Goldstone Modes of Superfluid Helium,” Phys. Rev. D, vol. 100, no. 11, p. 116 007, 2019. arXiv: 1907.10635 [hep-ph].

[42] A. Kamenev, Field Theory of Non-Equilibrium Systems. Cambridge University Press, 2011.

[43] P. Romatschke and U. Romatschke, Relativistic Fluid Dynamics In and Out of Equilibrium, ser. Cambridge Monographs on Mathematical Physics. Cambridge University Press, May 2019, isbn: 978-1-108-48368-1, 978-1-108-75002-8. arXiv: 1712.05815 [nucl-th].

[44] V. I. Ogievetsky, “Nonlinear realizations of internal and space-time symmetries,” in X-th winter school of theoretical physics in Karpacz, Poland. 1974.,

[45] E. A. Ivanov and V. I. Ogievetsky, “The Inverse Higgs Phenomenon in Nonlinear Realizations,” Teor. Mat. Fiz., vol. 25, pp. 164–177, 1975.

[46] I. Low and A. V. Manohar, “Spontaneously broken space-time symmetries and Gold- stone’s theorem,” Phys. Rev. Lett., vol. 88, p. 101 602, 2002. arXiv: hep-th/0110285.

[47] S. Weinberg, The quantum theory of fields. Vol. 2: Modern applications. Cambridge Uni- versity Press, 1996.

[48] L. V. Delacrétaz, S. Endlich, A. Monin, R. Penco, and F. Riva, “(Re-)Inventing the Relativistic Wheel: Gravity, Cosets, and Spinning Objects,” JHEP, vol. 11, p. 008, 2014. arXiv: 1405.7384 [hep-th].

[49] S. Endlich, A. Nicolis, and R. Penco, “Ultraviolet completion without symmetry restora- tion,” Phys. Rev. D, vol. 89, no. 6, p. 065 006, 2014. arXiv: 1311.6491 [hep-th].

[50] M. Greiter, F. Wilczek, and E. Witten, “Hydrodynamic Relations in Superconductiv- ity,” Mod. Phys. Lett. B, vol. 3, p. 903, 1989.

[51] D. T. Son, “Low-energy quantum effective action for relativistic superfluids,” Apr. 2002. arXiv: hep-ph/0204199.

153 [52] P. Glorioso and D. T. Son, “Effective field theory of magnetohydrodynamics from generalized global symmetries,” Nov. 2018. arXiv: 1811.04879 [hep-th].

[53] G. Goon, A. Joyce, and M. Trodden, “Spontaneously Broken Gauge Theories and the Coset Construction,” Phys. Rev. D, vol. 90, no. 2, p. 025 022, 2014. arXiv: 1405.5532 [hep-th].

[54] L. V. Delacrétaz, B. Goutéraux, S. A. Hartnoll, and A. Karlsson, “Theory of hydro- dynamic transport in fluctuating electronic charge density wave states,” Phys. Rev. B, vol. 96, no. 19, p. 195 128, 2017. arXiv: 1702.05104 [cond-mat.str-el].

[55] S. Grozdanov and N. Poovuttikul, “Generalized global symmetries in states with dy- namical defects: The case of the transverse sound in field theory and holography,” Phys. Rev. D, vol. 97, no. 10, p. 106 005, 2018. arXiv: 1801.03199 [hep-th].

[56] M. Baggioli and A. Buchel, “Holographic Viscoelastic Hydrodynamics,” JHEP, vol. 03, p. 146, 2019. arXiv: 1805.06756 [hep-th].

[57] J. Armas and A. Jain, “Viscoelastic hydrodynamics and holography,” JHEP, vol. 01, p. 126, 2020. arXiv: 1908.01175 [hep-th].

[58] P. C. Martin, O. Parodi, and P. S. Pershan, “Unified hydrodynamic theory for crystals, liquid crystals, and normal fluids,” Phys. Rev. A, vol. 6, pp. 2401–2420, 6 1972.

[59] L. P. Pitaevski˘ı, “SECOND SOUND IN SOLIDS,” Soviet Physics Uspekhi, vol. 11, no. 3, pp. 342–344, 1968.

[60] N. Ashcroft and N. Mermin, Solid State Physics. Philadelphia: Saunders College, 1976.

[61] D. Forster, Hydrodynamic Fluctuations, Broken Symmetry, And Correlation Functions. CRC Press, 1975.

[62] R. Holyst and A. Poniewierski, “On the elastic free energy for smectic-A liquid crystals,” Journal de Physique II, vol. 3, no. 2, pp. 177–182, 1993.

[63] D. Gaiotto, A. Kapustin, N. Seiberg, and B. Willett, “Generalized Global Symmetries,” JHEP, vol. 02, p. 172, 2015. arXiv: 1412.5148 [hep-th].

[64] J. Zhao, Z. Yan, M. Cheng, and Z. Y. Meng, “Higher-form symmetry breaking at Ising transitions,” Nov. 2020. arXiv: 2011.12543 [cond-mat.str-el].

[65] E. Lake, “Higher-form symmetries and spontaneous symmetry breaking,” Feb. 2018. arXiv: 1802.07747 [hep-th].

154 [66] Y. Hidaka, Y. Hirono, and R. Yokokura, “Counting Nambu-Goldstone modes of higher- form global symmetries,” Jul. 2020. arXiv: 2007.15901 [hep-th].

[67] D. M. Hofman and N. Iqbal, “Goldstone modes and photonization for higher form symmetries,” SciPost Phys., vol. 6, no. 1, p. 006, 2019. arXiv: 1802.09512 [hep-th].

[68] L. V. Delacrétaz, D. M. Hofman, and G. Mathys, “Superfluids as Higher-form Anoma- lies,” SciPost Phys., vol. 8, p. 047, 2020. arXiv: 1908.06977 [hep-th].

[69] D. Harlow and H. Ooguri, “Symmetries in quantum field theory and quantum gravity,” Oct. 2018. arXiv: 1810.05338 [hep-th].

[70] S. Grozdanov, D. M. Hofman, and N. Iqbal, “Generalized global symmetries and dissi- pative magnetohydrodynamics,” Phys. Rev. D, vol. 95, no. 9, p. 096 003, 2017. arXiv: 1610.07392 [hep-th].

[71] S. Grozdanov, A. Lucas, and N. Poovuttikul, “Holography and hydrodynamics with weakly broken symmetries,” Phys. Rev. D, vol. 99, no. 8, p. 086 012, 2019. arXiv: 1810. 10016 [hep-th].

[72] S. Grozdanov and N. Poovuttikul, “Generalised global symmetries in holography: mag- netohydrodynamic waves in a strongly interacting plasma,” JHEP, vol. 04, p. 141, 2019. arXiv: 1707.04182 [hep-th].

[73] J. Armas, J. Gath, A. Jain, and A. V. Pedersen, “Dissipative hydrodynamics with higher-form symmetry,” JHEP, vol. 05, p. 192, 2018. arXiv: 1803.00991 [hep-th].

[74] J. Armas and A. Jain, “Magnetohydrodynamics as superfluidity,” Phys. Rev. Lett., vol. 122, no. 14, p. 141 603, 2019. arXiv: 1808.01939 [hep-th].

[75] ——, “One-form superfluids \& magnetohydrodynamics,” JHEP, vol. 01, p. 041, 2020. arXiv: 1811.04913 [hep-th].

[76] N. Iqbal and N. Poovuttikul, “2-group global symmetries, hydrodynamics and holog- raphy,” Oct. 2020. arXiv: 2010.00320 [hep-th].

[77] J. Armas and A. Jain, “Hydrodynamics for charge density waves and their holographic duals,” Phys. Rev. D, vol. 101, no. 12, p. 121 901, 2020. arXiv: 2001.07357 [hep-th].

[78] B. Horn, A. Nicolis, and R. Penco, “Effective string theory for vortex lines in fluids and superfluids,” JHEP, vol. 10, p. 153, 2015. arXiv: 1507.05635 [hep-th].

[79] A. Cepellotti, G. Fugallo, L. Paulatto, M. Lazzeri, F. Mauri, and N. Marzari, “Phonon hydrodynamics in two-dimensional materials,” Nature Communications, vol. 6, no. 1, p. 6400, 2015.

155 [80] A. J. Beekman, J. Nissinen, K. Wu, K. Liu, R. Slager, Z. Nussinov, V. Cvetkovic, and J. Zaanen, “Dual gauge field theory of quantum liquid crystals in two dimensions,” Mar. 2016. arXiv: 1603.04254 [cond-mat].

[81] J. de Boer, M. P. Heller, and N. Pinzani-Fokeeva, “Holographic Schwinger-Keldysh effective field theories,” JHEP, vol. 05, p. 188, 2019. arXiv: 1812.06093 [hep-th].

[82] P. Glorioso, M. Crossley, and H. Liu, “A prescription for holographic Schwinger- Keldysh contour in non-equilibrium systems,” Dec. 2018. arXiv: 1812.08785 [hep-th].

156 Appendix A: Emergent gauge symmetries

In this section, we offer a somewhat pragmatic explanation for why we expect non- equilibrium effective actions to possess the gauge symmetries (2.15-2.16). It is pragmatic in the sense that it should convince the reader that these symmetries are necessary in ordinary situations, but it is by no means a derivation of the symmetries from first principles. In the ordinary construction of the hydrodynamic equations of motion it is usually sup- posed that the system exists in ‘local equilibrium.’ This essentially means that, ignoring broken Goldstones, the state of the system can be specified by the local inverse tempera- ture four-vector βµ(x) as well as the chemical potentials µA(x) corresponding to unbroken

symmetries TA. From the EFT perspective, we posit that the fluid manifold is a space on

M M which the fluid four-vector is fixed, namely β = (β0, 0, 0, 0) , where β0 is the equilibrium inverse temperature. Then, in the physical spacetime, we can define the inverse-temperature four-vectors—one for each leg of the SK contour—via the push-forwards

µ µ µ M ∂Xs ∂Xs β (x) ≡ β = β0 . (A.1) s ∂σM ∂σM

Additionally, we can identify the chemical potential with1

A A A Bs0 µs (x) = µ0 + √ , (A.2) −G00

A A where Bs0 is given in (3.30) and µ0 are the equilibrium chemical potentials of the unbroken symmetries. Working in the r, a-basis of (1.18), the r-type fields correspond to classical field

1It may not be immediately obvious that this combination of fields should be identified with the chemical potential. For some insight into why this should be so, see the discussion surrounding (2.40-2.42) and [3, 52].

157 configurations, whereas the a-type fields describe quantum and statistical fluctuations. As a result, we ought to identify the physical thermodynamic quantities βµ(x) and µA(x) with

1 βµ(x) = βµ(x) ≡ βµ(x) + βµ(x), r 2 1 2 (A.3) 1 µA(x) = µA(x) ≡ µA(x) + µA(x). r 2 1 2

If we wish to reproduce ordinary hydrodynamics, then we must require that the equations of motion only depend on the unbroken Goldstone fields as they appear in (A.3). To ensure that this is the case, we must require that the effective action be invariant under the symme- tries (2.15-2.16). Further, since we expect that the (classical) state of the system is specified

µ A by βr (x) and µr (x), these emergent symmetries cannot change the state of the system and therefore ought to be considered as mere gauge redundancies. It is worth mentioning that beyond these heuristic arguments, there has been some success in deriving chemical shift symmetries for unbroken U(1) Goldstones using the AdS/CFT duality from first principles [81, 82].

158 Appendix B: Stückelberg tricks and the Maurer-Cartan form

The purpose of this section is to include many of the important results of §3.6.1 for readers who do not wish to think about higher-form symmetries. We claimed in §2.2 that the correct way to construct covariant building-blocks for non- equilibrium effective actions was to construct two distinct Maurer-Cartan one-forms (3.30)— one for each leg of the SK contour—that transform under a single copy of the global symmetry group G. In this section, we will use a Stückelberg trick inspired by [2] to demonstrate that using two copies of the Maurer-Cartan form is the correct approach. We begin by introducing sources for the Noether currents corresponding to each symmetry generator of G. This amounts to introducing external gauge fields. Let

P¯m = unbroken translations,

TA = other unbroken generators, (B.1)

τα = broken generators, be the generators of G, where we now use m, n = 0, 1, 2, 3 to denote Lorentz indices and µ, ν to denote physical spacetime coordinate indices.1 As before, let H be the set of unbroken symmetries and H0 ⊂ H the subgroup generated by TA alone. The external sources are as follows:

m • Let εsµ(x) be the vierbeins, which can be thought of as the gauge fields correspond-

ing to unbroken translations P¯m. The metric tensors are then given by gsµν(x) =

m n εsµ(x)ηmnεsν(x).

1It is now necessary to distinguish between Lorentz indices m, n and physical spacetime coordinate indices µ, ν because the Stückelberg trick requires that we gauge all symmetries including Lorentz.

159 A • Let Asµ(x) ≡ Asµ(x)TA be the gauge fields (or spin connections if the Lorentz group is

involved) corresponding to the unbroken symmetries of H0 ⊂ H.

α • Let csµ(x) ≡ csµ(x)τα be the gauge fields (or spin connections) corresponding to the broken symmetries.

On each leg of the SK contour, we can combine these fields into a single object,

m ¯ α A θsµ(x) = iεsµ(x)Pm + icsµ(x)τα + iAsµ(x)TA, (B.2)

0 where the factors of i are included for later convenience. Now, letting U(t, t ; θsµ) for s = 1, 2 0 be the operator from t to t in the presence of external sources θsµ, the generating functional for the conserved currents is

W[θ1µ,θ2µ]  †  e = tr U(+∞, −∞; θ1µ)ρU (+∞, −∞; θ2µ) . (B.3)

Since θsµ couple to conserved currents, W[θ1µ, θ2µ] must be invariant under two independent copies of the gauge symmetries [2]; that is, for gauge parameters ζ1(x) and ζ2(x) we have

ζ1 ζ2 W[θ1µ, θ2µ] = W[θ1µ, θ2µ]. (B.4)

We can therefore ‘integrate in’ both the broken and unbroken Goldstone fields using the Stückelberg trick. In particular, define

 µ  ∂X ∂ α A −1 s iπ (σ)τα i (σ)TA ΘsM (σ) ≡ γ · θsµ(Xs(σ)) + γs, γs(σ) ≡ e s e s , (B.5) s ∂σM ∂σM where σM for M = 0, 1, 2, 3 are the fluid worldvolume coordinates. Now we can implicitly define the non-equilibrium effective action by

¹ W[θ1µ,θ2µ] µ α A iIEFT[Θ1M,Θ2M ] e ≡ DXs Dπs Ds e . (B.6)

160 m m A α Notice that if we remove the external source fields by fixing εsµ(x) = δµ and Asµ = csµ = 0, we find that ΘsM are nothing other than the Maurer-Cartan one-forms (3.30) defined on

m m each leg of the SK contour. Since εsµ(x) = δµ we can identify the Lorentz indices m, n with the physical spacetime coordinate indices µ, ν, allowing us to use P¯µ to refer to the unbroken translation generators as we did in (2.17). Moreover because these fields live on the SK contour, their values must match up in the infinite future, meaning that even though there were two copies of the gauge fields and gauge symmetries, there is only one copy of the global symmetry group G.

161 Appendix C: Explicit inverse Higgs computations

Since the thermal and unbroken inverse Higgs constraints are new, it may be helpful to see how they can be solved explicitly. This appendix explains how to algebraically solve the thermal and unbroken inverse Higgs constraints in the case of ordinary fluids, thereby removing the broken boost Goldstones ηi and the unbroken rotations Goldstones θi, respec- tively. But before we solve these new inverse Higgs constraints at full nonlinear order, it is first helpful to understand the motivations for imposing them. To do this, we will investigate these inverse Higgs constraints expanded to linear order in the fields.

µ µ µ Consider the case of a fluid without conserved charge. Let Xs (σ) = σ + εs (σ) and µν i 0i let the antisymmetric field θs (σ) represent all Lorentz Goldstones such that ηs = θs and i 1 i jk jk θs = 2  θs . Then at linear order in the fields, the vierbeins are given by

µ µ µ µ EsM = δM + ∂M εs − θsM + ··· , (C.1) where s = 1, 2 indicates on which leg of the SK contour the fields are defined. We are µ interested in setting to zero components of EsM that transform covariantly and that can give µ µ an algebraic relation between θsM and derivatives of εs . i To remove the broken boost Goldstones, notice that Es0 for i = 1, 2, 3 transform covari- antly under all symmetries and gauge symmetries and are given by

i i i Es0 = ∂0εs − θs0 + ··· . (C.2)

162 Thus setting the above expression to zero gives the linearized constraint

i i ηs = −∂0εs, (C.3)

i which is just the linearized version of (2.31). This allows the removal of the ηs-Goldstones. Removal of the rotational Goldstones is a bit trickier. There are no terms in a single

i vierbein that transform covariantly and, when set to zero, allow the removal of θs. However, we have two vierbeins—one for each leg of the SK contour—thus we consider the object

−1 M E1Mi(E2 )j , which transforms covariantly under all symmetries and gauge symmetries. Set- ting the antisymmetric part to zero gives the relation

−1 M −1 M 0 = E1Mi(E2 )j − E1M j(E2 )i = ∂j εai − ∂iεaj − 2θaji + ··· , (C.4)

i i i i j ij ij where εa ≡ ε1 − ε2 and θa ≡ θ1 − θ2 . Solving the above equation gives the constraint

ì ì θa = ∇ × εìa, (C.5)

which is the linearized version of (2.33). Then, acting with the (classical) dynamical KMS symmetry on both sides gives ì ì ∂0θr = ∇ × ∂0εìr, (C.6)

i 1 i i  i j 1 i j i j ì where εr ≡ 2 ε1 + ε2 and θr ≡ 2 (θ1 + θ2 ). Since θr enjoys a chemical shift-type gauge ì ì symmetry, at linear order, θr can only appear in the form ∂0θr . Thus, the constraints (C.5- ì C.6) are sufficient to remove the θs-Goldstones entirely from the (quadratic) effective action.

163 C.1 Thermal inverse Higgs

Recall that for ordinary fluids, the thermal inverse Higgs constraints are given by (2.30). Since R is invertible, these constraints give

µ t j ∂X i ∂X i ∂X i 0 = Λµ = Λt + Λj . (C.7) ∂σ0 ∂σ0 ∂σ0

i i ì Let λ be the 3×3 matrix with components λ j = Λj and let l be the 3-vector with components

i i l = Λt . Then we have

λ = (1 − ηˆ ⊗ ηˆ) + ηˆ ⊗ ηˆcosh η, lì = ηˆsinh η, (C.8)

p where η ≡ ηì · ηì and ηˆ ≡ ηì/η. Now equation (C.7) becomes

∂ Xì 0 − −1 · ì t = λ l, (C.9) ∂0X which simplifies to ∂ Xi ηi 0 − t = tanh η, (C.10) ∂0X η

as desired.

C.2 Unbroken inverse Higgs

Recall that for ordinary fluids, the unbroken inverse Higgs constraints are given by (2.32). Expanding out the vierbeins, we find that

ρ µ j i ν ∂Xs ν E = asM Rs j , asM ≡ Λsρ , (C.11) sM ∂σM

164 i i j iθ (σ)Ji i j where Rs = [e s ] . Then, we have that

−1 M −1 j −1 M (Es )i = (Rs )i (as )j , (C.12)

which gives us

−1 M l −1 k −1 M −1 k l (E1)iM (E2 ) j = (a1)M (R1)li(R2 )j (a2 )k ≡ (R2 )j Mk (R1)li, (C.13)

µ l −1 M l −1 ν ∂X1 l where Mk ≡ (a )k (a1)M = (Λ )k ν (Λ1)µ . Then unbroken inverse Higgs constraints 2 2 ∂X2 (2.32) yield the 3 × 3 matrix equation

T T T R2 · M · R1 = R1 · M · R2, (C.14)

T −1 T where we have used the fact that Rs = Rs . Rearranging this equation we have that M = −1 −1 (R1 · R2 )· M ·(R1 · R2 ), from which it is immediate that

T −1 T −1 T −1 2 M · M = (R1 · R2 )· M · M ·(R1 · R2 ) = (R1 · R2 · M) . (C.15)

√ −1 T Finally, we have that (R1 · R2 )· M = M · M, that is,

√ −1 T −1 R1 · R2 = M · M · M , (C.16)

as desired. Notice that the above expression, when linearized, gives (C.5), so it can be used to remove a-type rotational Goldstones. To remove the r-type rotational Goldstones beyond the linearized level requires a bit more work; we will carry out the computation in the next subsection.

165 C.3 More on unbroken inverse Higgs

Now we will see that the unbroken inverse Higgs constraints allow the removal of the r- √ T −1 type rotational Goldstones beyond linear order in the fields. First, define Ma ≡ M · M and suppose that under the (classical) dynamical KMS symmetries, we have

Ma → ΘMa − iβΘ∂0Mr, (C.17)

where Θ is a time-reversing symmetry of the UV theory.1 Then, by applying the (classical) dynamical KMS transformations to (C.16), we have

T T R1 · Ωr0 · R2 = ∂0Mr =⇒ Ωr0 = R1 · ∂0Mr · R2, (C.18)

1 where ΩrM ≡ 2 (Ω1M + Ω2M ) is the r-type spin connection. The above equation is our second unbroken inverse Higgs constraint and when linearized, gives (C.6). It can be checked that 0 in every invariant building-block, Rs can appear without σ -derivatives only in the form

T R1 · R2 . But then we can use (C.16) to remove this combination of rotational Goldstones.

Further, if Rs appears in an invariant building-block, the only other package it can come in is Ωr0, in which case we can use (C.18) to remove Ωr0. This expedient comes at the price of introducing more factors of Rs. It turns out that there are two possibilities:

• The additional factors of Rs are contracted in such a way that they cancel completely,

in which case we have no remaining factors of Rs.

T • The additional factors of Rs appear in the form R1 · R2 in which case we can use (C.16) to remove them.

Therefore, no matter what, the rotational Goldstones can be entirely removed from the invariant building-blocks using (C.16) and (C.18). It should be noted that the covariant

1 Think of (C.17) as a definition of Mr .

166 building-blocks may still have factors of Rs attached to them that cannot be removed with inverse Higgs-type constraints, but this is not an issue as only the invariant building-blocks are relevant for constructing EFTs.

167 Appendix D: Fluids and volume-preserving diffeomorphisms

It turns out that despite appearances, the effective actions for fluids with and without charge that we constructed with cosets, (2.37) and (2.42) respectively, are equivalent to those given in [9]. Additionally, the action for finite-temperature superfluids (2.50) that we constructed with cosets turns out to be equivalent to the action presented in [12]. In this section, we will see how these seemingly different actions are equivalent. For definiteness, consider the action (2.37), which is defined on the fluid worldvolume. Perform a change of coordinates so that it is now defined on the physical spacetime; then the dynamical degrees of freedom are the fields σM (x) and they enjoy the gauge symmetry

σM (x) → σM (x) + ξ M (σI(x)), (D.1) for arbitrary ξ M . As always, we use M, N = 0, 1, 2, 3 and I, J = 1, 2, 3. The resulting action defined on the physical spacetime is

¹ 4 SEFT = d x P(T), (D.2)

1 µ 0 where T ≡ u ∂µσ such that

µ µ J p µ  1 2 3 u ≡ , b ≡ −J Jµ, J ≡ ? dσ ∧ dσ ∧ dσ . (D.3) b

0 0  Jµ 0  µ Now integrate out σ . The equation of motion for σ is 0 = ∂µ b P (T) . Since ∂µJ ≡ 0 identically, this becomes   µ 1 0 0 = J ∂µ P (T) . (D.4) b

1 2 It turns out that T = −1/G00.

168 By the definition of J µ, we have that

1 P0(T) = f (σI), (D.5) b for some arbitrary function f (σI). Using the diffeomorphism gauge symmetry (D.1), we can gauge-fix f (σI) = 1. With this gauge-fixing condition, (D.5) can be solved algebraically for T as a function of b, that is T = T(b). Plugging this solution back into the expression for the effective action (D.2) gives ¹ 4 SEFT = d x F(b), (D.6) where F(b) ≡ P(T(b)). But this is precisely the action given in [9]. Notice that now the action only depends on the three fields σI, which enjoy a volume-preserving diffeomorphism gauge symmetry ∂gI σI → gI(σJ), det = 1. (D.7) ∂σJ

By almost identical procedures, it can be shown that (2.42) is equivalent to the effec- tive action describing charged fluids given in [9] and that (2.50) is equivalent to the action describing finite-temperature superfluids given in [12].

169