<<

Non-Fermi as ersatz Fermi liquids: general constraints on compressible

Dominic V. Else,1 Ryan Thorngren,2 and T. Senthil1 1Department of , Massachusetts Institute of Technology, Cambridge, MA 02139, USA 2Center for Mathematical Sciences and Applications, Harvard University, Cambridge, MA 02138 A system with charge conservation and lattice translation symmetry has a well-defined filling ν, which is a real number representing the average charge per unit cell. We show that if ν is fractional (i.e. not an integer), this imposes very strong constraints on the low- theory of the system and give a framework to understand such constraints in great generality, vastly generalizing the Luttinger and Lieb-Schultz-Mattis theorems. The most powerful constraint comes about if ν is continuously tunable (i.e. the system is charge-compressible), in which case we show that the low-energy theory must have a very large emergent symmetry group – larger than any compact Lie group. An example is the Fermi surface of a Fermi , where the charge at every point on the Fermi surface is conserved. We expect that in many, if not all, cases, even exotic non-Fermi liquids will have the same emergent symmetry group as a Fermi liquid, even though they could have very different dynamics. We call a system with this property an ersatz Fermi liquid. We show that ersatz Fermi liquids share a number of properties in common with Fermi liquids, including Luttinger’s theorem (which is thus extended to a large class of non-Fermi liquids) and periodic “quantum oscillations” in the response to an applied magnetic field. We also establish versions of Luttinger’s theorem for the composite Fermi liquid in quantum Hall systems and for spinon Fermi surfaces in Mott insulators. Our work makes connection between filling constraints and the theory of symmetry-protected topological (SPT) phases, in particular through the concept of “’t Hooft anomalies”.

CONTENTS VII. Extension to spinful systems 16

I. Introduction 2 VIII. Fermi surfaces coupled to dynamical gauge fields 17 II. Kinematic properties of the IR theory 3 A. HLR theory of composite Fermi liquid 17 A. Example: in 1 spatial B. Son theory of composite Fermi liquid 19 dimension 4 C. Spinon Fermi surface 20

III. Filling constraints 5 IX. Impossibility of Fermi arcs and other A. One spatial dimension 5 constraints on the Fermi surface 21 1. Example: Luttinger liquid in one spatial A. Fermi arcs 21 dimension 6 B. More general constraints 22 B. Two spatial dimensions 6 C. General space dimension 6 X. The role of fractionalization 23

IV. Consequences of the filling constraints for XI. Discussion: relation to “violations of compressible states 6 Luttinger’s theorem” 25

V. The kinematic properties of Fermi liquids 7 XII. Conclusion and outlook 26 A. Emergent symmetry group 7 Acknowledgments 26 B. The homomorphism GUV → GIR 8 C. The ’t Hooft anomaly 8 A. Filling does not correspond to the boundary of D. Filling constraint and Luttinger’s theorem 9 an SPT 26 E. Extension to higher dimensions 10 arXiv:2007.07896v4 [cond-mat.str-el] 6 Apr 2021 B. Translation symmetry fractionalization on a 2π VI. Ersatz Fermi liquids and their phenomenology 11 flux in 2-D systems at fractional filling 27 A. General properties of EFLs 11 B. Response to electric fields 12 C. Filling constraints in general dimension from C. Quantum oscillations 13 the topological action of SPTs 27 1. Two spatial dimensions 13 2. Two spatial dimensions: multiple Fermi D. Projective representation of loop group on surfaces 14 monopole from ’t Hooft anomaly 28 3. Three spatial dimensions 15 D. Anomalous Hall effect 15 E. Projective representation of loop group in a E. Chiral magnetic effect 15 Fermi liquid 28 2

F. Proof of Theorem 1 29 U(1) symmetries and the emergent symmetries of the IR theory; and (c) a property of the IR theory called its G. Bounds on Filling Denominators for Compact “ ’t Hooft anomaly”. Some connections between filling GIR 30 and ’t Hooft anomalies have previously been explored in Refs. [15–20]. H. Asymptotic analysis for quantum oscillations 32 Our work has important implications for the study of References 33 “non-Fermi liquids”, which are systems that are metallic down to zero temperature but for which the IR physics cannot be described by Fermi liquid theory. Motivated by I. INTRODUCTION the results on filling constraints just mentioned, we will introduce the concept of an ersatz Fermi liquid (EFL) In , systems with prescribed as a general framework to understand non-Fermi liquids. microscopic degrees of freedom (usually ) can An EFL is a system which has the same kinematic prop- exhibit varied and exotic emergent behavior at low ener- erties in the deep infrared as a Fermi liquid, though gies. In general, it is extremely difficult, either analyti- it might have very different dynamical properties. By cally or numerically, to predict the nature of the emergent “kinematic” properties we mean properties that relate low-energy behavior [as described by an “IR (infra-red) to the structure of the Hilbert space that describes the theory”] from the properties of the microscopic degrees of ground state and the low-energy excitations, as opposed freedom. For this reason, it is invaluable to have general to “dynamical” properties which relate to the Hamilto- results that constrain the nature of the IR theory, given nian that acts in this Hilbert space. More precisely, the microscopic properties of the system. kinematic properties comprise properties (a), (b) and (c) One such result is the Lieb-Schultz-Mattis theorem [1], described in the previous paragraph and in more detail which states (in the formulation of interest to us here [2]) in the next section. Further, it will will turn out that the that in a system in one spatial dimension with a con- kinematic properties we will discuss have a strong topo- served U(1) charge and discrete translation symmetry, if logical flavor, so in a suitably vague sense we can say that the average charge per unit cell (which we call the “fill- an EFL is a system that is “topologically equivalent” to ing”) is not an integer, then either the IR theory is gap- a Fermi liquid. (However, since the dynamical properties less or else it spontaneously breaks one of the symmetries. of an EFL can be sharply different to that of a Fermi This result was later generalized to higher dimensions by liquid, it will generally not be the case that an EFL can Oshikawa [3] and Hastings [4]; in higher dimensions there be continuously deformed into a Fermi liquid.) is also the possibility that a system at fractional filling can be gapped but with non-trivial [5]. We will show that many (though by no means all) A related result is Luttinger’s theorem [6], which states well-known aspects of Fermi liquid phenomenology are, that if the IR theory is a Fermi liquid, then the volume in fact, purely kinematic in nature, and therefore apply enclosed by the Fermi surface (modulo the volume of the equally well to any EFL. In particular, we will show that Brillouin Zone) is determined entirely by the fractional any EFL has a Fermi surface that hosts long-lived ex- part of the filling, and in particular is independent of the citations (though these excitations may not be Landau interaction strength. Originally proven perturbatively by ). In fact, our approach leads to a very Luttinger, the result was later established through a non- general perspective on what it means for a system to perturbative argument (but still assuming that the IR have a Fermi surface. Furthermore, we will find that theory is describable by Fermi liquid theory) by Oshikawa this Fermi surface must obey Luttinger’s theorem (or a [7]. A generalized version of Luttinger’s theorem even generalization thereof, analogously to the “fractionalized holds in a class of phases - known as Fractionalized Fermi Fermi liquids” mentioned above); and that, if the Fermi Liquids [8, 9] - which are distinct from conventional Fermi surface geometry is such that a Fermi liquid with that liquids. In such phases a gapless Fermi liquid co-exists geometry would exhibit quantum oscillations in the de- with non-trivial topological order. The generalization to pendence of physical properties on magnetic field, then Luttinger’s theorem can then exactly be determined from any EFL with the same Fermi surface geometry is also the interplay of the symmetry with the topological order expected to display quantum oscillations with the same [9–11]. periodicity. These results and others [12–14] raise the question: what is the most general statement that one can make Since none of the kinematic properties that define an about the relation between the microscopic filling and EFL require that the system be weakly coupled or have properties of the IR theory? In the present work, we will a description in terms of quasiparticles, we expect a wide answer this question, showing that the microscopic fill- variety of exotic non-Fermi liquid phenomena to be re- ing is always completely determined by a few properties alizable within the class of EFLs. In fact, we will argue of the IR theory, namely: (a) its emergent symmetries; based on the general theory of filling constraints that (b) the relation between the microscopic translation and any IR theory which describes a compressible , i.e. 3 the filling can be continuously tuned1 must have a very fied by some classes of non-Fermi liquids. However, as non-trivial emergent symmetry group, larger than any the is destroyed in the IR in such problems, compact Lie group. Such a property is indeed satisfied (even to all orders) should be used by EFLs (due to the infinitely many conserved quanti- with caution. In a non-perturbative context, despite the ties associated with the Fermi surface); whether it could formal existence[24] of a Luttinger-Ward functional, the be satisfied in a different way that leads to fundamen- conventional proof of Luttinger’s theorem fails. This is tally different kinematic properties is an important open dramatically illustrated by the fractionalized Fermi liquid problem. phases. It is thus desirable to have a more general non- At the very least, however, it is clear that a number of perturbative argument for Luttinger’s theorem in this non-Fermi liquid metals can be fruitfully discussed from context, which we provide in this paper. the perspective developed in this paper. The simplest are non-Fermi liquid metals that arise when a Fermi surface is coupled to a critical boson that represents a fluctu- II. KINEMATIC PROPERTIES OF THE IR ating order parameter. Within the standard framework THEORY for such quantum critical points, they will be seen to be EFLs. A closely related system is a Fermi surface cou- Among all the features of of the IR theory describing pled to a gapless U(1) gauge field, which arises in the the low-energy properties of a system, there are certain theory of composite Fermi liquid metals in the quantum ones that we will refer to as “kinematic”, and these are Hall regime, and in some insulating quantum liq- the subject of this section. uids with Fermi surfaces of emergent electrically neutral Before describing the first kinematic property, let us quasiparticles. In these states we will find that, unlike note that the microscopic system will have a group of a strict EFL, the kinematic properties of the IR theories symmetries GUV. We will specifically be interested in can differ from those of a Fermi liquid (indeed, unlike systems where this includes a global U(1) corresponding a Fermi liquid, these are not states in which the micro- to charge conservation, and translation symmetries (pos- scopic density can be continuously tuned) though still sibly on a lattice). closely related. The main point will still stand, though, The first kinematic property of the IR theory is the set that the kinematic properties of these IR theories provide of emergent symmetries. We thus introduce the group a powerful framework to thinking about their universal 2 GIR of emergent symmetries of the IR theory. GIR is, behavior. In particular we will show that in such systems in general, not the same as the microscopic symmetry a version of Luttinger’s theorem is still satisfied. group GUV. Nevertheless, each element g ∈ GUV of the As a final application of our results, we will examine microscopic symmetry group gets mapped into an ele- the possibility for systems to exhibit disconnected Fermi ment ϕ(g) ∈ GIR of the emergent symmetry group, such arcs instead of a closed Fermi surface. We will find that, that ϕ(g1)ϕ(g2) = ϕ(g1g2) for all g1, g2 ∈ G. There assuming the translational symmetry is unbroken, such could, of course, be elements of the emergent symmetry a scenario is inconsistent with the IR theory being an group GIR that do not correspond to any microscopic EFL, except when the system exists on the boundary of symmetry. Also, a microscopic symmetry g ∈ GUV could a gapless bulk (such as a Weyl ). This pro- act trivially in the IR theory, in which case ϕ(g) = 1 vides strong evidence for the impossibility of Fermi arcs. (the identity element in GIR); an example of the latter In fact, we obtain a stronger constraint: in an EFL, the case would be for systems with a charge gap, in which Fermi surface must enclose a volume in the Brillouin zone; this volume is the generalization to EFLs of the “ sea” in a weakly interacting system. Before proceeding, let us make a final technical re- 2 For most of the paper we will focus on the most familiar kind mark. We note that previous work[21–23] generalized of emergent symmetry; what has retroactively been renamed “0- the original perturbative proof of Luttinger’s theorem form symmetry”, i.e. symmetries that act everywhere in space- to situations where there are a number of and time at once. Recently, the importance of so-called “higher-form boson fields with U(1) symmetries (either global or symmetries” has started to be understood [25]. Loosely speak- gauge). This proof relies on the existence of a Luttinger- ing the restriction to the ordinary 0-form symmetries means that there are no emergent “fractionalized excitations”. We say a Ward functional of the exact Green’s functions of the point-like excitation is “fractionalized” if it cannot be created /bosons from which the self- can be ex- by a point-like creation operator, but rather is created at the tracted by functional derivatives. These works lead to endpoints of an open string by an operator supported on the the expectation that Luttinger’s theorem will be satis- string. There are similar definitions for higher-dimensional exci- tations. Fractionalized excitations emerge, for instance, in phases where there is topological order and/or deconfined emergent gauge fields. For such systems, there will be emergent higher- form symmetries that will play an important role, but in the 1 A perhaps more standard definition of “compressible” would be interest of simplifying the exposition we will defer discussion of that the partial derivative of the particle number with respcet to this point to later, and for the moment we assume that all the the is nonzero, however we will not use this emergent symmetries are 0-form symmetries. We will eventually definition in this paper. return to this point in Section X. 4 case the microscopic U(1) symmetry acts trivially in the in the presence of background gauge fields. Thus, the IR. All these statements can be expressed in a compact classification of ’t Hooft anomalies in d spatial dimen- mathematical way by saying that ϕ defines a group ho- sions is precisely equivalent to the classification of SPT 3 momorphism from GUV to GIR that need not be injective phases in d + 1 spatial dimensions . Such classification or surjective. has been explored at great depth, from a variety of per- Note that for many of the arguments in this paper, we spectives, ranging from physical considerations to very will specifically want GIR to represent the internal sym- formal ones[26, 27, 31, 37–61] The aspects of SPT phases metries of the system (that is, the symmetries which do that we will need to use in this paper, however, will be not move space-time points around). In general, the IR simple enough, at least if we want to understand filling in theory will also have “trivial” emergent space-time sym- systems of low spatial dimension d ≤ 2, that a reader un- metries such as continuous translation symmetry, which familiar with this literature should still be able to follow we do not include in GIR. It is important to note that the our paper. microscopic translation symmetry in general will not map into these trivial translation symmetries, but rather into the internal symmetry GIR. One way to think about this A. Example: Luttinger liquid in 1 spatial is that since there is a spatial rescaling transformation dimension associated with passing to the IR theory, a microscopic translation symmetry in fact has trivial translation ac- Let us illustrate the above general considerations in tion in the IR limit. An alternative perspective is that the case of a system of spinless electrons in a lattice in we imagine that a microscopic translation symmetry acts one spatial dimension. Thus, the microscopic symmetry like a product of an internal symmetry and a “trivial” group GUV is comprised of a U(1) symmetry generated translation symmetry, and we only worry about the in- by the total electron number Qˆ, and a Z symmetry gen- ternal part in defining the map ϕ. erated by the lattice translation operator T. The other property of the IR theory that will be per- Let us now assume that the IR theory of the electrons tinent is the extent to which the full emergent symmetry is a Luttinger liquid. Thus, the low-energy physics takes group GIR can be ‘naturally’ realized in some realization place at the two Fermi points at momenta kL and kR. (not necessarily the original microscopic lattice model) Excitations with close to kL are left-movers, of the IR theory. This property is formalized by the and those with momentum close to kR are right-movers. concept of the ’t Hooft anomaly. Such an anomaly in At low energies, the numbers NˆL and NˆR of left- and d + 1 space-time dimensions is an obstruction to UV- right-movers are separately conserved. Therefore, they regularizing the theory on a lattice in d spatial dimen- generate the emergent symmetry group GIR = U(1)L × sions with the full emergent symmetry group GIR real- U(1)R. ized as an “on-site” microscopic symmetry [26, 27]. A Next, we need to specify how the microscopic symme- powerful alternative but formal characterization of an ’t try acts on the IR theory, which we can do by expressing Hooft anomaly [28, 29] is that the conservation law corre- the generators of the microscopic symmetry in terms of sponding to the GIR symmetry is broken upon coupling the generators of the emergent symmetry. Indeed, we to a background gauge field for the symmetry GIR. We have will consider to the ’t Hooft anomaly of the emergent ˆ ˆ ˆ symmetry GIR to also be a ‘kinematic’ property of the Q ∼ NL + NR (1) IR theory. ˆ ˆ T ∼ exp(−i[kLNL + kRNR])., (2) ’t Hooft anomalies are also closely related to the theory of symmetry-protected topological (SPT) phases [30–32]. where the tilde “∼” refers to an equivalent action on the These are gapped phases of matter in a system with an IR theory. unbroken global symmetry G with the following prop- The emergent symmetry group GIR in a Luttinger liq- erty: the ground state cannot be continuously deformed uid has the well known axial anomaly. This is an example into a trivial product-state ground state while preserving of the ’t Hooft anomaly mentioned in the previous subsec- the symmetry without closing the gap if the symmetry tion. The signature is that, if we turn on an electric field G is preserved, but can be if the symmetry G is lifted. E for the U(1) symmetry generated by NˆL + NˆR [which Examples of SPT phases include the celebrated topolog- is equivalent to a microscopic electric field, by Eq. (1)], ical insulators [33] and the Haldane of the spin-1 then left- and right-moving charges are no longer sepa- antiferromagnetic chain in d = 1 [34–36]. µ rately conserved; instead, if we let jL , µ = 0, 1 be the The connection between ’t Hooft anomalies and SPTs current density for the left-moving charge, and similarly is that a G SPT phase in d + 1 spatial dimensions must have a non-trivial boundary theory, and the boundary theory carries a ’t Hooft anomaly. We can say that the ’t Hooft anomaly of the boundary theory is “canceled” 3 There can also be ’t Hooft anomalies associated with “invertible” by inflow from the bulk, in the sense that the conser- topological phases which do not require any symmetry to protect vation laws of the bulk+boundary system are preserved them. 5

µ for jR , we have theory is completely trivial) can have any integer filling. The fundamental observation we will make is that the E ∂ j µ = − , microscopic filling is completely fixed by the kinematic µ L 2π properties of the IR theory, i.e. the emergent symmetry E group and the ’t Hooft anomaly, along with the mapping ∂ j µ = . (3) µ R 2π from the microscopic symmetry group into the emergent symmetry group. One can easily understand this equation in the case of a This is a rare example of a precise relation between a non-interacting Fermi . In that case, in the absence microscopic quantity (the filling) and properties of the IR of electric field the electrons occupy single-particle states theory. Such relations are extremely useful, given that labeled by momentum k. But applying an electric field it is usually very difficult to determine the IR theory pointing to the right causes an overall flow of electrons ˙ from the microscopic Hamiltonian, either analytically or in momentum space according to k = E. This causes a numerically. charge excess to accumulate at kR and a corresponding The reason why such a UV-IR correspondence is pos- charge deficit at kL. sible in this case is because the UV and IR are linked As we mentioned above, there is always a bulk- through the homomorphism ϕ that implements the mi- boundary correspondence that relates a ’t Hooft anomaly croscopic symmetry inside the emergent IR symmetry. for a symmetry G in d spatial dimensions to an SPT in Naively, we can imagine arguing as follows. The lattice d + 1 spatial dimensions whose boundary theory carries filling is defined in systems that have at least a micro- the ’t Hooft anomaly. In this case, we have d = 1 and scopic U(1) symmetry (charge conservation), and a lat- G = U(1)×U(1). The corresponding SPT phase in d = 2 tice translation symmetry d (where d is the space di- is realized by a “quantum spin Hall” state corresponding Z mension). So we set the microscopic symmetry GUV = to putting spin-up electrons in a quantum Hall state with d U(1) × Z . This GUV is embedded into the symmetries quantized Hall conductance σxy = 1 and spin-down elec- of the IR theory, and so we can talk about the ’t Hooft trons in a quantum Hall state with σxy = −1. Here the anomaly of the IR theory thought of as a GUV-symmetric U(1) × U(1) symmetries correspond to the separate con- theory (which can be computed from the G ’t Hooft servation of spin-up and spin-down electrons. If we now IR anomaly in light of the homomorphism ϕ : GUV → GIR). consider a system with boundary, and apply an electric Then we invoke “UV-IR anomaly matching” to relate the field parallel to the boundary, due to the Hall conduc- filling ν to this G ’t Hooft anomaly. tance this will generate a current of spin-up electrons UV Unfortunately, this argument is not quite right. In incident onto the boundary, and a current of spin-down fact, fractional filling cannot correspond to a nontrivial ’t electrons with opposite sign. This precisely accounts for Hooft anomaly for G in the usual sense, because there for the charge non-conservation in the boundary theory UV are no candidate SPTs with G = U(1) × d symmetry due to the ’t Hooft anomaly. UV Z which could constitute the “bulk”, for which a system Finally, let us remark that a useful way to think about with fractional filling could be the boundary, as we ex- SPT phases (and hence ’t Hooft anomalies) is in terms plain in Appendix A. Indeed, a system with fractional of topological terms describing the response of the SPT filling seems perfectly well-defined on its own without phase to background gauge fields. For example, consider any need for a higher-dimensional bulk. the quantum spin Hall state described above. We can Instead, we will give the correct version of the argu- theoretically couple to background gauge fields of the two ment below, in various spatial dimensions. The careful U(1) symmetries, which in the one-dimensional Luttinger reader might notice that they still appear reminiscent of liquid were interpreted as the conservation of left- and “anomaly matching”, albeit for a ’t Hooft anomaly that is right-movers; hence, we denote the gauge fields by AL trivial with respect to the microscopic symmetries. How and AR. The response of the quantum spin Hall state is exactly one could make such a notion precise, we leave then described by a Chern-Simons action on (2 + 1)-D as an open question. Nevertheless, the arguments are spacetime: self-contained and hold without any need for such an in- 1 Z terpretation. S[A] = (AR ∧ dAR − AL ∧ dAL). (4) 4π

A. One spatial dimension III. FILLING CONSTRAINTS In one spatial dimension, we imagine putting the sys- Now we will turn to the question of how to under- tem on a circle and then very slowly threading 2π flux stand constraints on the IR theory resulting from the of the microscopic U(1) symmetry through the circle microscopic filling, i.e. the average charge per unit cell, [3, 7, 62], By a standard argument, this transforms the which is a real number ν. Only the fractional part of ground state into a low-lying excited state with a differ- ν (i.e. ν mod 1), should be expected to be detectible in ent momentum; if we label states by their eigenvalues of the IR theory, because an atomic (whose IR the translation operator T, i.e. T|ψi = e−ip|ψi, then the 6

−ip −i(p+2πν) momentum gets shifted by e → e . translations Tx and Ty obey in the limit as the 2π flux Meanwhile, since the microscopic U(1) symmetry, gen- becomes infinitely spread out spatially) the magnetic al- erated by Qˆ, corresponds to a U(1) symmetry of the IR gebra ˆ theory, whose generator we call QIR, we can also imag- −1 −1 2πiν ine performing the 2π flux insertion in the IR theory. TxTyTx Ty = e . (8) Now consider the IR symmetry τ = ϕ( ) ∈ G corre- T IR Heuristically, this is clear because a 2π flux sees a back- sponding to microscopic translation T. In the IR theory, the ground state can get transformed into a low-lying ground charge density as an effective magnetic field. We excited state with a different eigenvalue of τ; that is, give some more careful arguments in Appendix B. This the eigenvalue of τ gets shifted according to exp(−iθ) → result is also closely connected to the translational sym- exp (−i[θ + α]) for some α [defined mod 2π]. One can metry fractionalization of a monopole in an insulator in argue that α depends only on the ’t Hooft anomaly of 3 spatial dimensions in the presence of polarization, as discussed in Ref. [20]. GIR and on the choice of τ and QˆIR; see for instance the ˆ Meanwhile, in the IR theory we can consider 2π flux example below. We write α = α(QIR|τ) (the dependence ˆ on the ’t Hooft anomaly is kept implicit). configurations of the IR symmetry generated by QIR that Now, the key point is that the processes described in corresponds to the microscopic U(1) symmetry. The the two paragraphs above are in fact the same process, homomorphism ϕ maps Tx and Ty into some elements just described in two different ways. Therefore, we must τx, τy ∈ GIR. In the presence of a ’t Hooft anomaly, such equate 2π fluxes indeed can carry a projective representation, i.e.

α(QˆIR|τ) ν = (mod 1). (5) V (τ )V (τ )V (τ )−1V (τ )−1 = eiα, (9) 2π x y x y where V (g) denotes the action of a group element g ∈ 1. Example: Luttinger liquid in one spatial dimension GIR on the 2π flux, and α = α(QˆIR|τx, τy) [defined mod 2π] depends on QˆIR, τx and τy, and on the ’t Hooft We can apply the general framework described above anomaly of GIR. Therefore, we must identify to the particular example of a Luttinger liquid in one ˆ spatial dimension, as discussed in Section II A. In this α(QIR|τx, τy) ν = (mod 1). (10) case QˆIR and τ are defined by the right-hand sides of 2π Eqs. (1) and (2). The natural example to consider to illustrate this con- Threading the flux of Qˆ generates an electric field IR straint will be a Fermi liquid. However, by contrast to of Qˆ by Faraday’s law, and then, from the anomaly IR the one-dimensional case, the ’t Hooft anomaly of a Fermi equation Eq. (3) we see that the 2π flux threading creates liquid in two spatial dimensions has not previously been −1 charge of U(1) and +1 charge of U(1) . Therefore, L R discussed. This is the subject of Section V. in light of Eq. (2), the momentum transforms according to

e−ip → e−ipe−i(kR−kL). (6) C. General space dimension

Then, from Eq. (5) we find that The formulation of filling constraints discussed above can be generalized to arbitrary space dimension d. The 1 ˆ ν = (k − k ) (mod 1), (7) idea is to generalize the functions α(QIR|τ) (in d = 1) and 2π R L α(QˆIR|τx, τy) (in d = 2) to a function α(QˆIR|τ1, ··· , τd) which is nothing other than Luttinger’s theorem for a that determines the filling in general spatial dimension d. Luttinger liquid in one spatial dimension. A convenient way to express this function is in terms of the “topological action” that describes the SPT phase in d + 1 spatial dimensions which cancels the GIR ’t Hooft B. Two spatial dimensions anomaly in d spatial dimensions by inflow on the bound- ary. We give the details in Appendix C. Here the idea is to consider a “2π flux” of the micro- scopic U(1) symmetry; that is, a very weak background magnetic field spread out over some very large region, IV. CONSEQUENCES OF THE FILLING such that the total flux is 2π. We can then consider CONSTRAINTS FOR COMPRESSIBLE STATES how such a 2π flux transforms under translation symme- try. We claim that, in the presence of fractional filling ν, The systematic theory of filling constraints described such 2π fluxes exhibit translational symmetry fractional- in the previous section has a very important corollary. ization; that is, acting on a 2π flux, the lattice x and y We want to consider IR theories which can exist at 7 generic filling; that is, they are not pinned to a par- system, we see that it has a total filling nν. Since at ticular filling but instead the filling can be continuously this filling the IR theory has no anomaly, it follows that tuned. In other words, the IR theory represents a “com- νn = p with p an integer, which is essentially the claim of pressible” state. What we will show is that in this case, Theorem 1. In d = 3, this simple argument does not work for spatial dimensions ≥ 2, the emergent symmetry GIR because there are SPTs in d + 1 = 4 space dimensions cannot be a compact Lie group. In one dimension, re- such that there is no finite n for which n copies become call that the Luttinger liquid example discussed in Sec- trivial. (This is also why the argument does not work for tion II A achieves generic filling with only a compact Lie d = 1). However, Theorem 1 still holds for any d ≥ 2 as group emergent symmetry GIR = U(1)×U(1). Note that we show in Appendices F and G. compact Lie groups include finite groups as special cases, since we do not require the Lie group to be connected; also, in this paper when we refer to Lie groups, we will V. THE KINEMATIC PROPERTIES OF FERMI always assume they are finite-dimensional. LIQUIDS We wish to emphasize here that as we stated previ- ously, in this paper when we refer to the emergent sym- In this section, we will return to a familiar IR theory: metry group GIR, we are referring specifically to the in- a Fermi liquid in two spatial dimensions (we will briefly ternal symmetries of the IR theory. While translation discuss higher dimensions as well) and analyze its kine- symmetry is never a compact group, we do not know matic properties, in the language introduced previously. of any examples of theories that can sensibly arise in We will see how Fermi liquids, by virtue of having an condensed matter systems, for which the internal sym- emergent symmetry group that is “larger” than a com- metry is a non-compact Lie group, and this may in fact pact Lie group, are able to evade the theorem of the be impossible. Therefore, our results suggest that the previous section and exist at generic filling. emergent symmetry for IR theories that exist at generic filling must be an infinite-dimensional group4. Indeed, this is the case for Fermi liquids, as we describe in the A. Emergent symmetry group next section. Our main result is the following: The first step is to identify the emergent symmetry Theorem 1. Suppose the emergent symmetry group of group GIR. We invoke the following well-known property the IR theory is some compact Lie group GIR. Then for of Fermi liquids: non-forward scattering terms are irrele- any spatial dimension d ≥ 2, the filling ν is constrained vant in the RG sense at low energies, so the quasiparticle to be an integer multiple of 1/NGIR , for some finite in- number at each point on the Fermi surface is separately teger NGIR that depends only on the group GIR and the conserved at low energies. Thus, Fermi liquids have a dimension. very large emergent symmetry group [63]. Roughly, we can say that G = “U(1)∞”. However, let us be a bit The proof is straightforward based on the framework IR more precise about how one approaches the “∞”. of filling constraints discussed in the previous section. We parameterize the Fermi surface by a continuous pa- We give the details, and reveal what determines NGIR , in Appendix F, and also give a more formal point of view rameter θ (we do not require that θ literally represents in Appendix G. a geometrical angle), which is periodic, i.e. it lives on a Let us mention a simple way to understand this result circle. Imagine that we place an IR cutoff on the sys- for d = 2. As mentioned in the introduction, the ’t Hooft tem (that is, place it in finite volume), which since the anomaly of a theory in d spatial dimensions implies that Fermi surface exists in momentum space, corresponds to the theory can be realized as the boundary of an SPT a short-distance cutoff on θ, i.e. θ now takes discrete val- phase protected by the same symmetry G in d+1-space ues. To each such θ value, we associate a U(1) emergent IR ˆ dimensions. Now suppose that the SPT is such that n symmetry generated by an integer-valued operator Nθ. copies of it is trivial, with n a finite positive integer. For Hence, a general symmetry operator will take the form SPTs protected by a compact Lie group GIR in space ! dimension d + 1 = 3, this is known to be always true. X exp −i f Nˆ (11) For the d-dimensional boundary theory of interest, this θ θ means that for n copies, there is no ’t Hooft anomaly. θ If now we consider n copies of the microscopic lattice where we identify fθ ∼ fθ + 2π. Now we send the IR cutoff (the spatial volume) to in- finity, which corresponds to sending the spacing between discrete θ values to zero. What we want is to consider 4 We remind the reader again of the restriction that for the mo- ment we assume that all the emergent symmetries are 0-form symmetry operators that do not depend too sensitively symmetries, which we re-examine in Section X. In fact, theorem on the precise way in which the short-distance cutoff in 1 continues to hold in the presence of finite higher form symme- θ gets sent to zero. In order to achieve this, we require tries. that, in this continuum limit, the fθ parameters become 8 smooth functions f(θ). Therefore, in the limit, the emer- C. The ’t Hooft anomaly gent symmetries are in one-to-one correspondence with smooth functions from the circle into U(1). The group Now we are in a position to discuss the ’t Hooft of all such functions is called the loop group [64] of U(1) anomaly for the emergent LU(1) symmetry. As usual, and we denote it by LU(1). Hence we conclude that the ’t Hooft anomaly can be understood by coupling to GIR = LU(1). a background gauge field for the symmetry. But first, We emphasize that the group structure of LU(1) dif- we must ask, what is a gauge field for an LU(1) symme- fers somewhat from naive conceptions of what “U(1)∞” try? Since, roughly speaking, an LU(1) symmetry means would mean; in particular, LU(1) has only one U(1) sub- there is a U(1) symmetry for each point on the circle, we group, whose elements correspond to taking f(θ) to be can naively say that a gauge field for an LU(1) symmetry a constant function. Physically, this is because only the should be a space-time vector field Aµ(θ) for each point θ total charge on the Fermi surface is quantized to be an on the circle, with gauge transformations parameterized integer; there is no well-defined concept of the (quan- by a scalar field λ(θ), and gauge transformation tized) charge at a single point on the Fermi surface, only of the linear charge density with respect to θ. Accord- Aµ(θ) → Aµ(θ) + ∂µλ(θ) (15) ingly, we can represent the loop group LU(1) formally by (where the derivative is respect to spacetime, not θ). introducing a density operatorn ˆ(θ) such that the num- Moreover, we require that A (θ) and λ(θ) be smooth ber of quasiparticles between θ and θ + dθ is measured µ functions of θ. byn ˆ(θ)dθ. Technically,n ˆ(θ) is not really an operator in However, there is in fact an additional ingredient that itself, but an operator-valued distribution which should is required, that is unique to loop groups. To see be integrated against a test function. The elements of this, we can appeal to the quasiparticle picture of the the emergent symmetry group can be expressed as Fermi liquid. The spatial components Ai(θ) describe the  Z  Aharanov-Bohm phase (which can be interpreted as a exp −i f(θ)ˆn(θ) , (12) Berry phase) picked up as a spatially localized quasipar- ticle, localized near position θ on the Fermi surface, is transported in space. But we can also keep a quasiparti- where f(θ) is any smooth function of θ. Note that, be- cle fixed in real space, and transport it along the Fermi cause we identify f(θ) ∼ f(θ)+2π, we are allowed to con- surface in momentum space. Therefore, the gauge field sider functions f with non-trivial winding number around needs an additional component Aθ to describe the Berry 1 R the circle, such that 2π ∂θf(θ)dθ is any integer. phase associated with such a process. We emphasize, though, that even going beyond Fermi liquids, Aθ rep- resents an intrinsic part of what it means to couple to a gauge field for an LU(1) symmetry, regardless of any B. The homomorphism GUV → GIR quasiparticle picture. We will return to this point in Sec- tion VI. Now it should be clear how to embed the microscopic The Aθ component of the gauge field transforms under symmetries into the emergent symmetry group GIR = gauge transformations as A → A + ∂ λ. Therefore, if LU(1). Indeed, if Qˆ is the generator of the microscopic θ θ θ we now combine the vector field Aµ with Aθ, we obtain a U(1) symmetry, we have vector field A on the D+1-dimensional manifold M ×S1, Z where M is the space-time manifold, D = dimM is the Qˆ ∼ q nˆ(θ)dθ, (13) space-time dimension, and S1 is the circle on which the θ variable lives. Moreover, A transforms under gauge transformations precisely as would a U(1) gauge field on where the integer q is the charge of a Landau quasipar- M × S1, Hence we arrive at our conclusion: An LU(1) ticle. Of course, for a Fermi liquid of electrons, q = 1, gauge field on M is equivalent to a U(1) gauge field on but in principle one can imagine Fermi liquid-like states M × S1. From this point on, we will denote this com- where the quasiparticles carry a different charge. (For bined vector field as Aµ, taking the convention that in- example, the quasiparticles could be bound states of an dices such as µ vary both over space-time directions and odd number of electrons). the θ direction. Meanwhile, if kx(θ) and ky(θ) represent the compo- Now we can discuss ’t Hooft anomalies. One way to nents of the lattice momentum of the point on the Fermi characterize an ’t Hooft anomaly in two spatial dimen- surface parameterized by θ, then we have sions is in terms of the topological term describing the re- sponse of the corresponding SPT in three spatial dimen-   Z sions to background gauge fields. [As usual, the relation Tα ∼ exp −i kα(θ)ˆn(θ)dθ , (14) is that on a (3+1)-D spacetime M+ whose boundary is a (2+1)-D manifold M, the SPT on M+ gives rise to the ’t where α = x, y, and Tx, Ty are the lattice translation Hooft anomaly on the space-time M by anomaly inflow.] operators. But from the above discussion, a topological term for 9

LU(1) gauge fields on a (3+1)-D spacetime M+ is equiv- has nonzero scattering rate for quasiparticles in the inte- alent to a topological term for U(1) gauge fields on the rior of the Fermi surface, but we can, theoretically, imag- 1 (3+1+1)-D manifold M+ × S . Hence, we conclude that ine an extension of the quasiparticle Hamiltonian from the appropriate topological term is the 5D Chern-Simons the vicinity of the Fermi surface to its interior in a man- action ner that preserves conservation of quasiparticle number Z at every k-point. This is a familiar construction in Fermi m liquid theory. (The point is that the physics on the Fermi S[A] = 2 A ∧ dA ∧ dA, (16) 1 24π M+×S surface will ultimately not depend on the precise form of the Hamiltonian in the interior.) where the level m is quantized to be an integer5. Below, Then by similar arguments to before, a gauge field for 2 we will show that such a topological term, with m = the LT U(1) symmetry is equivalent to a U(1) gauge field ±1, indeed reproduces many known properties of spinless 2 on M × T . Since this is also a 5-dimensional manifold, Fermi liquids. (Note that the sign of m is only defined we can write a similar Chern-Simons term to Eq. (16). relative to a choice of orientation for the Fermi surface, Specifically, we write because redefining θ → −θ sends m → −m). From the topological term Eq. (16) we can determine m Z S[A] = 2 A ∧ dA ∧ dA, (18) the anomaly equation on the boundary by computing the 24π M×D δS current j = δA and then considering the current incident onto the boundary. We find that, on the boundary, the where D ⊆ T2 is the volume “occupied by electrons”, continuity equation is violated according to that is, the volume enclosed by the Fermi surface6. This gives rise to the same anomaly equation Eq. (17) on the µ m λστκ boundary ∂(M × D) = M × ∂D, where ∂D is the Fermi ∂µj =  (∂λAσ)(∂τ Aκ). (17) 8π2 surface. Note that in this interpretation, the components A ,A ,A of A are the usual ones while the components Here the current j depends both on space-time coordi- 0 x y µ A ,A are k-space gauge fields. Then the 5 dimen- nates and on θ. Its space-time components describe the kx ky sional manifold has the interpretation of being “-time current of the charge at position θ on the space” (that is, the space where points are labeled by Fermi surface, i.e. of the symmetry generated byn ˆ(θ). position and momentum) plus time, and A can be con- However, recalling that indices are supposed to vary over µ sidered a gauge field in phase space. A term similar to the θ direction as well as space-time directions, we have Eq. (18) has previously appeared in Ref. [37]. to introduce the component jθ = δS , which describes δAθ flow of charge along the Fermi surface. An example of a θ case where j 6= 0 is a Fermi liquid in a magnetic field, D. Filling constraint and Luttinger’s theorem described in the next subsection. Finally, let us mention that there is an alternative pic- ture to understand inflow of the ’t Hooft anomaly that We consider a theory in two spatial dimensions whose is sometimes helpful. Instead of considering a system ’t Hooft anomaly is canceled by inflow from Eq. (16). Ac- cording to the discussion of Section III B, we first need to with LU(1) symmetry in a (3+1)-D space-time M+ with 2 determine how the emergent symmetry LU(1) gets rep- boundary, we can consider a system with LT U(1) sym- resented projectively in such a theory in the presence of metry in a (2+1)-D space-time M without boundary. 2 a 2π flux (of the microscopic charge U(1) symmetry). Here LT U(1) is the group of smooth maps from the Bril- This is something that can be derived from the ’t Hooft 2 louin zone (thought of as a torus T ) into U(1). In other anomaly that was characterized in the previous subsec- words, we imagine that the charge is conserved not just tion. We show in Appendix D that this leads to a pro- at each point on the Fermi surface, but also at every k- jective representation on a 2π flux described by the com- point in the whole Brillouin zone. This is the case, for mutation relations example, in a non-interacting . In an interact- mq ing Fermi liquid, the physical Hamiltonian presumably [ˆn(θ), nˆ(θ0)] = −i δ0(θ − θ0), (19) 2π wheren ˆ(θ) are the operators introduced in Section V A, and δ0 is the derivative of the Dirac delta function. 5 To see the quantization, observe that the total charge carried by the SPT, under the U(1) subgroup of LU(1) corresponding to setting the function f(θ) in Eq. (12) to be independent of θ, is given by the integral (over a fixed time-slice) 6 By a particle-hole transformation which cannot modify the Z Z Z physics, the “occupied volume” could equivalently be taken to 0 δS m j = = dA ∧ dA = mC[A], be the complement Dc of D in the Brillouin zone. Observe, how- δA 8π2 0 ever, that if we replaced D with Dc in Eq. (18), the only change where C[A] is the second Chern number and is quantized to be in the anomaly equation Eq. (17) on the Fermi surface would an integer. Hence, m must be an integer in order for the total be an additional minus sign, which as we mentioned earlier just charge of the SPT to always be an integer. corresponds to a choice of orientation of the Fermi surface. 10

Note that Eq. (19) is precisely the so-called Kac- (with a possible integer multiplicative factor mq) for any Moody algebra satisfied by the local density operator of a IR theory that has the same emergent symmetry as a chiral fermion in one spatial dimension [such as appears, Fermi liquid. for example, at the boundary of a integer quantum Hall It is interesting to reconsider Eq. (19) and its rela- state in two spatial dimensions]. Indeed, this relationship tion with the 5D Chern-Simons term from the viewpoint is not a coincidence; one can see roughly how it comes that the 5-dimensional manifold can be thought of as 4- about in the case of a Fermi liquid at the level of the dimensional phase space together with the time direction semiclassical theory of electron transport. Suppose, for (see the last two paragraphs of Section V C). In the pres- simplicity, that we switch off the interactions so that the ence of a static 2π-strength magnetic flux in the x, y com- Fermi liquid becomes a non-interacting Fermi gas. In the ponents of Aµ (with the corresponding components Ax presence of a spatially-dependent magnetic field B(x), and Ay independent of kx,ky,t), the 5D Chern-Simons the semiclassical equations of motion take the form [65] term Eq. (18) reduces to a 3D Chern-Simons term for the remaining components A = (A ,A ,A ), assum- dk I 0 kx ky = −qB(x) × v(k) (20) ing that they are independent of x and y: dt dx dk m Z = v(k) − Ω(k) × , (21) S = d3x A ∂ A (25) dt dt 3D 4π IJK I J K where Ω(k) is the Berry curvature of the Bloch states, ∂E(k) where kx and ky are integrated over the volume D, i.e. v(k) = ∂k , E(k) is the dispersion relation, and we have over the interior of the Fermi surface. omitted terms beyond linear order in the magnetic field Now the claim is that on the Fermi surface, there is a strength. In particular, Eq. (20) implies that electrons chiral mode carrying the Kac-Moody algebra develop a circulation in momentum space along contours of constant energy, in particular along the Fermi sur- m [ˆn(θ), nˆ(θ0)] = −i δ0(θ − θ0), (26) face. In two spatial dimensions the Fermi surface is one- 2π dimensional and the circulation along the Fermi surface is unidirectional (set by the sign of the magnetic field), In particular since a 2π flux of the microscopic U(1) corre- which indeed resembles a chiral fermion. sponds to a 2πq flux of A by Eq. (13), we recover Eq. (19). This intuitive argument, however, does not fix the co- In the phase space interpretation, Eq. (25) describes an efficient of the right-hand side of Eq. (19). In Appendix integer quantum Hall effect in momentum space in the E we give a more careful derivation of Eq. (19) from interior of the Fermi surface. In other words we think the semiclassical theory of electron transport, confirming of the rigid interior of the Fermi surface as hosting an that m = ±1 for a spinless Fermi liquid. Related expres- integer quantum Hall state in momentum space when we sions have previously been derived in Refs. [66, 67]. apply a 2π flux in real space. The Fermi surface is the From Eq. (19) we can compute the projective repre- boundary in momentum space of the interior, and hence sentation of the translation symmetry in light of the em- hosts a chiral edge state. bedding Eq. (14). We find

 mq Z dk (θ)  −1 −1 = exp i k (θ) y dθ (22) TxTyTx Ty 2π x dθ E. Extension to higher dimensions  V  = exp imq F , (23) 2π The description of the anomaly extends straightfor- wardly to higher dimensional Fermi liquids. Indeed, if M where VF is the volume in momentum space enclosed is the (d + 1)-D space-time manifold, and F is a (d − 1)- by the Fermi surface. (Here we have chosen a particular dimensional manifold parameterizing the Fermi surface, convention to define the orientation of the Fermi surface). which is a codimension-1 surface in the d-dimensional Hence, from Eq. (10), we conclude that Brillouin zone, then the emergent symmetry group is LF U(1), i.e. the group of smooth maps from F to U(1). VF Including the components of the Berry connection on the ν = mq (mod 1), (24) (2π)2 Fermi surface promotes the LF U(1) gauge field to a U(1) gauge field on the (2d)-dimensional manifold M × F . which (if we set m = q = 1) is precisely Luttinger’s theo- Then we can write down a topological term on a (d+2)- rem for a spinless Fermi liquid in two spatial dimensions. D spacetime M+ describing the SPT whose inflow gen- We wish to emphasize, however, that, in general, Lut- erates the anomaly of the Fermi liquid, given by the tinger’s theorem Eq. (24) follows directly from Eq. (19), (2d + 1)-dimensional Chern-Simons action: which in turn follows directly from the ’t Hooft anomaly. It was not necessary to assume anything about the dy- m Z namical properties of the Fermi liquid, e.g. the existence S[A] = A ∧ (dA ∧ · · · ∧ dA) . (27) (d + 1)!(2π)d M+×F | {z } of quasiparticles. Thus, Luttinger’s theorem also holds d times 11

VI. ERSATZ FERMI LIQUIDS AND THEIR then implies that the linear charge densityn ˆ(θ) at each PHENOMENOLOGY point of the Fermi surface is conserved at the fixed point. (There is, potentially, a dangerous inter-patch BCS cou- As we showed previously (Section IV), any IR theory pling in the pairing channel that could destroy these con- that can exist at generic filling must have a very large servation laws. For the Fermi surface coupled to a gauge symmetry group, “larger” than any compact Lie group. field, a weak pairing interaction is irrelevant at the IR In the simplest cases, this emergent symmetry group will fixed point[80] while in the Ising nematic quantum criti- be the same as that of a Fermi liquid, i.e. GIR = LU(1). cal point it is relevant. In the latter case the non-Fermi We refer to an IR theory with this emergent symmetry liquid metallic fixed point is preempted by the supercon- as an ersatz Fermi liquid (EFL). We note that EFLs rep- ducting instability. Our discussion then applies at a scale resent a class of theories; there could be many different above this instability.) EFLs with different dynamical properties. Here we will We also expect that more complex quantum critical examine the properties that all EFLs have in common. In points associated with the death of a Fermi surface can particular, as a Fermi liquid is an example of an EFL, this be subsumed under the umbrella of EFLs. Such critical section will provide a fresh perspective on many aspects points have been argued[81] to possess a critical Fermi of Fermi liquid phenomenology, showing that they arise surface even in the absence of Landau quasiparticles. As directly from the emergent symmetry and its ’t Hooft we will see below, the EFL description, if it indeed applies anomaly without any need to invoke the detailed dynam- to such quantum critical points, enables inferring many ical properties of a Fermi liquid. of their general properties. Going beyond Fermi liquids, an important example of states which we expect will be EFLs are associated with quantum critical points in metals which are not A. General properties of EFLs tied to a particular electron density. As a concrete ex- ample consider a putative associ- Let us return to a general EFL. Then, translation sym- 7 ated with the onset of Ising nematic order in a metal in metry must embed into GIR somehow: that is, we have d = 2. There is no particular electron density at which this transition will happen in any given system. Indeed,  Z  as microscopic parameters are changed continuously, we Tα ∼ exp −i kα(θ)ˆn(θ)dθ . (28) expect that the electron density at the transition will also change continuously without a change of universality for some U(1)-valued functions k (θ). The values of class. The universal critical properties of this transition α kα(θ) can be interpreted as momenta in the Brillouin are described by a theory of electrons near the Fermi sur- zone, so this defines a codimension-1 surface in the face coupled to the fluctuating order parameter modes. Brillouin zone. We can take this to be the definition (A model with similar structure also describes insulat- of a Fermi surface in a general EFL. Of course, this ing phases with a spinon Fermi “Fermi surface” may or may not have any signature in, surface coupled to a dynamical U(1) gauge field). In say, angle-resolved photoemission spectroscopy (ARPES) these models the resulting IR fixed point is not a Fermi measurements, and if it does the signature might be dif- liquid[68–79]. Nevertheless, from our point of view, these ferent in character from that of a Fermi liquid. metallic quantum critical points are expected to be EFLs. Nevertheless, the fact thatn ˆ(θ) is a conserved opera- (We will discuss neutral Fermi surfaces in insulators sep- tor for every θ tells us that, at any point θ on the Fermi arately in Sec. VIII C.) surface, there are infinitely long-lived excitations which We may understand why such critical points should be are forbidden from scattering away from that point. To EFLs as follows. In previous papers [71, 74–76] the ulti- see this, first observe that sincen ˆ(θ) is conserved, exci- mate IR fixed point was accessed through a ‘patch con- tations can be labeled by their corresponding eigenvalue struction’ which begins by by dividing the Fermi surface n(θ). An excitation at point θ on the Fermi surface is into small patches. It was then argued that the important characterized by coupling of the fermions within a single patch is to bo- son fluctuations with momentum tangential to the local n(θ0) = Nδ(θ0 − θ), (29) Fermi surface. This enables treating the full system by focusing on a pair of antipodal patches and ignoring their for some N. One can show that N is quantized to an coupling to other such pairs of patches. The patch width integer, because by the definition of LU(1), the operator is taken to zero at the end. In this scheme, the number of fermions within each patch is conserved. The assumption Z that this patch description captures the IR fixed point Nˆ = nˆ(θ)dθ (30)

must have integer eigenvalues, since exp(−2πiNˆ) is the 7 This is associated with spontaneously breaking C4 lattice rota- identity operator. Henceforth, we will refer to such ex- tion symmetry to C2. citations as Fermi surface quanta. From Eq. (28) and 12

Eq. (29), we see that a single Fermi surface quantum car- nˆ(θ) to be no longer conserved in the presence of the ries momentum k(θ). Fermi surface quanta are the gen- electric field. Specializing for simplicity to a Fermi liquid eralization of Landau quasiparticles to a general ersatz in two spatial dimensions, the total charge on a segment Fermi liquid. Their dynamics, however, could be very [θ, θ + dθ] of the Fermi surface gets shifted according to different from Landau quasiparticles. Moreover, apart from the fact that their number is quantized, the Fermi dky(θ) nˆ(θ)dθ → nˆ(θ)dθ + qaxEt 2 (32) surface quanta need not have particularly “quasiparticle- (2π) /(LxLy) like” properties. Nevertheless, the Berry’s phase of a spatially localized Fermi surface quantum as it is moved where dky(θ) = ky(θ + dθ) − ky(θ), and Lx and Ly are in space or over the Fermi surface is still a well-defined the linear dimensions of the system (normalized by ax quantity, which supplies the general interpretation of the and ay respectively, the unit cell dimensions) in the x and y direction. We have to divide by the denominator gauge field Aµ (including the Aθ component) discussed in Section V C. in the second term in Eq. (32) to take into account the Next, we can consider how the microscopic U(1) sym- density of single-particle states in momentum space in a metry embeds into LU(1). The quantization of charge, finite-size system. i(2π)Qˆ ˆ We will now show that Eq. (32) indeed follows from i.e. the requirement that e = 1 (where Q is the mi- the ’t Hooft anomaly. In order to do that, we will want croscopic charge operator), constrains the embedding to to reformulate Eq. (32) in a way that removes the explicit be of the form dependence on Ly. We write Eq. (32) as Qˆ ∼ qN,ˆ (31) d axLxLy dky(θ) nˆ(θ) = η(θ; Ly) := qE (33) for some integer q. We can interpret q as the charge of a dt (2π)2 dθ single Fermi surface quantum. Next we identify the difference Finally, we can consider the ’t Hooft anomaly of the

LU(1) symmetry. There is not much freedom, since as ∆η(θ) := η(θ; Ly + 1) − η(θ; Ly) (34) we saw in Section V the ’t Hooft anomalies are just a L dk (θ) classified by the level m of the 5D Chern Simons term x x y = qE 2 (35) Eq. (16) which is quantized to be an integer. In sum- (2π) dθ mary, once we have fixed the shape of the Fermi surface, with the density shift rate associated with applying a uni- the kinematic properties of an ersatz Fermi liquid (in the form electric field in the x direction in the presence of a sense of Section II) are captured by the integers q and “flux of y translation” symmetry around the y direction. m. (Note that there is a freedom that sends q → −q and To compute this shift from the ’t Hooft anomaly, we m → −m simultaneously by redefining the generators of start with the anomaly equation Eq. (17) and consider a LU(1); moreover, as we mentioned earlier, choosing the configuration where reverse orientation of the Fermi surface sends m → −m while leaving q fixed). Ax = Et, (36) Next we observe the arguments of Section V D can be Ay = −ky(θ)/Ly, (37) applied in any EFL. Therefore, we immediately conclude that the Fermi surface in any EFL satisfies Luttinger’s At = 0 (38) theorem in the form Eq. (24). Aθ is independent of x, y, t (39) In the remainder of this section we will consider various aspects of Fermi liquid phenomenology and argue that The motivation for our choice of Ay is that it ensures they hold equally well in any EFL. that the LU(1) flux around the y direction is given by  Z  exp −i ky(θ)ˆn(θ)dθ , (40) B. Response to electric fields which coincides with the expression Eq. (14) for the A property of Fermi liquids is that if a uniform electric translation symmetry operator Ty. Substituting into field is applied, the system responds in essentially the Eq. (17) and integrating over x and y gives same way as it would in a non-interacting Fermi gas, mq dk (θ) which is to say that the momenta of quasiparticles get ∂ n(θ) + ∂ J θ = E y a L , (41) shifted (assuming the electric field is in the x direction) t θ (2π)2 dθ x x according to kx → kx + axEt, where ax is the unit cell θ R θ R t size in the x direction8. This causes the density operators where J = j dxdy, and we can identify j dxdy with the expectation value n(θ) = hnˆ(θ)i. Hence, provided θ that ∂θJ = 0, we recover Eq. (35) if we set m = 1. θ Our assumption that ∂θJ = 0, even in the presence of 8 This appears because we have normalized our momenta to be a background electric field, requires a bit more explana- dimensionless. tion. J θ represents a charge circulation along the Fermi 13 surface. As we mentioned in Section V D, in a Fermi liq- same periodicity of the quantum oscillations. We will uid a magnetic field induces a chiral flow of quasiparticles not, however, make any statement about the amplitude along the Fermi surface, in which case Jθ ∝ Bn(θ) and of the quantum oscillations, and it remains possible that generally ∂θJθ 6= 0. So we need to address why the same this amplitude and its dependence on parameters such thing could not happen with an electric field. Of course, as temperature will still allow for Fermi liquids and non- in a Fermi liquid, one can easily convince oneself that Fermi liquids to be distinguished. it does not, but we want an argument that holds more Our task is complicated by the fact that the quantum generally in any EFL. oscillations are non-perturbative in the magnetic field, In a quantized formulation of the IR theory, the current and therefore, it is not clear that one expects them to operator Jˆθ is defined by be directly describable in terms of the deep IR theory. Instead, we will find it necessary to make appeal to a UV Z ˆ δH 2 completion. Jˆθ = d x (42) δAθ where Hˆ is the Hamiltonian of the IR theory. Then we 1. Two spatial dimensions know that, in the absence of an electromagnetic field, Jˆθ must satisfy (as an operator identity) the conservation Let us first consider the case of a system in two spatial law dimensions which microscopically has continuous trans- lation symmetry, with a microscopic charge density ρ. In d nˆ(θ, t) + ∂ Jˆ (θ, t) = 0 (43) such a case, in the presence of a magnetic field the x and y dt θ θ translation generators Px and Py fail to commute, but we On the other hand, in the absence of an external electro- can define a “magnetic unit cell” of volume bx × by, such magnetic field, by assumptionn ˆ(θ) should be conserved that the flux per magnetic unit cell is 2π and the discrete for each θ, since it is the generator of the emergent sym- translations Tx = exp(−ibxPx) and Ty = exp(−ibyPy) metry. Therefore, in the absence of electromagnetic field, do commute. Then we can define the magnetic filling the current operator identically satisfies νM = (bxby)ρ = 2πρ/B, which is the charge per mag- netic unit cell. Then we can apply all the usual results

∂θJˆθ = 0. (44) on filling with respect to this discrete translation symme- try; in particular, if νM is not an integer then the Lieb- Now observe that because the electric field E enters into Schultz-Mattis-Oshikawa-Hastings theorem [1–4] forbids the Hamiltonian “temporally”, i.e. through the time- the system from having a trivial (i.e. not topologically derivative of A, it is not possible for Jˆθ defined by ordered) gapped ground state, whereas such a state is Eq. (42) to depend on E because the Hamiltonian is de- permitted for integer νM . Moreover, if we assume that fined on a single time slice of space-time. Therefore, Jˆθ the ground state at a given νM is itself an EFL with a continues to satisfy Eq. (44) even with an applied elec- single Fermi surface, then Luttinger’s theorem for EFLs implies that at integer νM the Fermi volume must fill all tric field. On the other hand, it is possible for Jˆθ to depend on the applied magnetic field, as the Fermi liquid of the Brillouin zone corresponding to the magnetic unit example demonstrates. cell, i.e. the Fermi surface becomes degenerate, presum- ably leading to an instability. Most generally, the point is that since νM mod 1 always reflects properties of the C. Quantum oscillations IR theory according to our general framework of filling constraints, the nature of the IR theory must vary with ν mod 1. Recall that Fermi liquids display “quantum oscilla- M These considerations motivate our assumption that, in tions” when a weak magnetic field B is applied; that is, general, observable properties of the ground state vary various physical properties are periodic in 1/B [82]. In periodically with ν , with period 1 (The periodicity the case that the physical property under consideration is M refers to the behavior for ν  1, and where ν varies resistivity, for example, this is known as the Shubnikov- M M over an interval ∆ν  ν . On longer scales there de Haas effect. (What we mean by “weak magnetic field” M M will be some envelope function governing the amplitude is that the magnetic flux per unit cell should be much less of the oscillations.) This behavior is well known in the than 1. To observe the oscillations at finite temperature Fermi liquid case, where integer ν (non-integer ν ) T , it is also necessary that the cyclotron energy E should M M c corresponds to fully filled (partially filled) Landau lev- satisfy E T , which for T > 0 also places a lower bound c & els respectively. Our periodicity assumption immediately on the magnetic fields for which the oscillations are ob- implies that, if we keep ρ fixed and vary B, then the pe- servable). riodicity with respect to 1/B is It is sometimes stated that observing quantum oscilla- tions is evidence for a quasiparticle description, i.e. of a ∆(1/B) = 1/(2πρ). (45) Fermi liquid. Here, however, we will show that there are very general reasons to expect any EFL to display the Next we can consider a system in two spatial dimen- 14

charge in this UV completion must be ρ = νo/(axay). Then with ρ so defined we again find that the periodicity of quantum oscillations is given by Eq. (45).

2. Two spatial dimensions: multiple Fermi surfaces

As a warm-up to going to three spatial dimensions, we will extend the results of the previous section to the case where we have several disconnected components of the Fermi surface, instead of just one. For simplicity we will FIG. 1. A Fermi liquid with a Fermi surface that wraps non- consider the case of two components, although the argu- trivially around the Brillouin torus does not exhibit quantum ments can easily be generalized. In that case, the emer- oscillations. gent symmetry is LF U(1), i.e. the group of smooth maps from F into U(1), where F = S1 t S1, i.e. the disjoint sions that microscopically only has discrete translation union of two circles. We can represent this by writing the (λ) symmetry. Here we recall that, even for Fermi liquids, generators of the emergent symmetry asn ˆ (θ), where 1 there is a condition on the Fermi surface in order to ob- θ ∈ S and λ ∈ {1, 2} labels the two components of the serve quantum oscillations. For a Fermi surface as de- Fermi surface. Similarly we write the momentum of the picted in Figure 1, no quantum oscillations will be ob- Fermi surface as k(λ)(θ). We assume that each compo- served; physically one can think of this as coming from nent satisfies the condition of no winding described ear- the fact that the semiclassical orbits of this Fermi sur- lier, which allows us to define the momentum operator face in a magnetic field are not closed in position space. The general condition for a Fermi liquid to display quan- Z ˆ X −1 (λ) (λ) tum oscillations is that each connected component of the Pα = aα kα (θ)ˆn (θ)dθ. (48) Fermi surface should have trivial winding number on the λ Brillouin torus. Let us therefore examine what we can say about a gen- The crucial point is that the fact that there are two con- eral EFL that satisfies this condition. To avoid some sub- nected components of the Fermi surface allows us to de- tleties in the argument, we will assume that q = 1. The fine two different U(1) emergent symmetries, generated key point is that when the Fermi surface does not have by Qˆ(1) and Qˆ(2), with nontrivial winding number, then there is a consistent way Z to define the Fermi momentum k(θ) without any mod 2π Qˆ(λ) = nˆ(λ)(θ)dθ, (49) ambiguity. Then we can define the filling that is relevant for quantum oscillations as In the same spirit as the argument of the previous section, m Z dk (θ) therefore, we will postulate that there is a UV completion ν = V , V = k (θ) y dθ, (46) o (2π)2 F F x dθ with both continuous translation symmetry and Qˆ(1) and Qˆ(2) realized microscopically. Defining ρ(1) and ρ(2) to be where VF is the volume enclosed by the Fermi surface, the corresponding charge densities, by similar arguments and this formula holds without any mod 1 equivalence. as earlier we have a generalized Luttinger’s theorem Note that ν is not necessarily the same as the micro- o Z scopic filling ν, although by Luttinger’s theorem it must (λ) m (λ) (λ) (λ) (λ) ρ = 2 VF , VF = kx (θ)ky (θ)dθ. differ from it by an integer. Moreover, we can define (2π) axay emergent symmetry generators (50) Z Now, in the presence of magnetic field, if we define mag- ˆ −1 (λ) (λ) Pα = aα kα(θ)ˆn(θ), (47) netic fillings νM = 2πρ /B, we expect that any observ- (1) (2) able property should be periodic in each of νM and νM with ax, ay the dimensions of the unit cell, such that with period 1. In other words, any observable property exp(−iaαPα) ∼ Tα. O can be expressed as This leads us to believe (although we will not give a (1) (2) rigorous proof) that, even though the actual microscopic O = f(νM , νM ), (51) theory may have had only a discrete translation sym- metry, there exists a UV completion of the IR theory where f(ν1, ν2) is some function such that f(ν1 +1, ν2) = with continuous translation symmetry whose action on f(ν1, ν2 + 1) = f(ν1, ν2). This implies that if we vary (λ) the IR theory is generated by Pˆα. Assuming that this B while keeping ρ fixed, and assuming that the ratio is the case, invoking the continuous translations version ρ(1)/ρ(2) is an irrational number, then O varies quasiperi- of Luttinger’s theorem shows that the density of U(1) odically in 1/B with base frequencies 2πρ(1) and 2πρ(2). 15

3. Three spatial dimensions the anomalous Hall conductance that, by contrast to the Berry phase contribution above, involve an integral over Now we can consider a 3D system. By dimensional gauge invariant quantities defined in local patches of the reduction, we can consider a 3D system as a 2D system; Fermi surface. but the Fermi surface of the 2D system will consist of in- Next we show that the ‘t Hooft anomaly of an EFL di- finitely many components corresponding to taking slices rectly implies the Berry phase contribution to the anoma- lous Hall effect. To that end we start from Eq. (18) through the 3D Fermi surface with fixed kz. By general- izing the previous discussion to N components, and then and consider a configuration where the gauge field A taking the limit as N → ∞, we find that any observable has a flux ΦB through the interior of the Fermi surface property should be expressible as in the (kx, ky) plane, and take the corresponding gauge

field components (Akx ,Aky ) to be independent of (t, x, y). O = F[ν⊥], (52) Similar to the discussion surrounding Eq. (25), we now find that the 5D Chern-Simons term reduces (assuming where F is some functional of ν⊥, and ν⊥ is a function that A0,Ax,Ay are independent of kx and ky) to: into R/Z (i.e. the real line with ν ∼ ν + 1 identified), mΦ Z defined by S = B dtdxdy  A ∂ A (57) AH 8π2 IJK I J K 1 ν (k ) = 2πρ (k ) mod 1 (53) ⊥ z ⊥ z B for the I, J, K = (0, x, y). This then directly corresponds ΦB mV⊥(kz) to a contribution to the Hall response σxy = 2π . We ρ⊥(kz) = 2 . (54) can write this directly in terms of the boundary theory (2π) axay R at the Fermi surface as ΦB = dθAθ. In the free fermion where V⊥(kz) is the two-dimensional area enclosed by the case this is exactly the Berry phase contribution to the intersection of the Fermi surface with a plane of fixed kz. anomalous Hall effect discussed above. In a general EFL, We show in Appendix H that this implies, so long as we should regard this as an unavoidable contribution due the functional F is sufficiently regular, that the depen- to the ’t Hooft anomaly of the IR theory. Like in the in- dence of O on B at small B is dominated by the extremal teracting Fermi liquid, the full measured anomalous Hall Fermi surface cross-sections, and we obtain that O varies effect may include other contributions that are ‘local’ on periodically or quasiperiodically with 1/B, with base fre- the Fermi surface. Note that Eq. (57) has the structure ∗ ∗ of a 3D Chern-Simons term but with a coefficient that is quencies 2πρ⊥(kz ), kz ∈ Σ, where Σ is the set of solutions d 1 to V⊥(kz) = 0. This is the same result as for a Fermi not quantized to be an integer multiples of 4π . This is al- dkz liquid. lowed here because the boundary gauge fields are coupled to the gapless modes associated with the Fermi surface. The full boundary action that includes both the Fermi D. Anomalous Hall effect surface modes and the unquantized Chern-Simons action of Eq. (57) will be properly gauge invariant (including The anomalous Hall effect refers to a Hall conductance under large gauge transformations). that exists in a time reversal broken system in zero exter- nal magnetic field. In general there can be many different effects that contribute to this effect. For a clean free elec- E. Chiral magnetic effect tron system with a Fermi surface, there is an interesting contribution that arises due to the net Berry curvature The chiral magnetic effect refers to a phenomenon in a ΦB of the filled Fermi sea. In d = 2, we have three-dimensional Fermi liquid where if the Chern num- ber of the Berry curvature on the two-dimensional Fermi ΦB surface is nonzero (which, in a non-interacting Fermi gas, σ = (55) xy 2π would occur when the Fermi surface encloses a Weyl point), then the total charge becomes nonconserved in This contribution can alternately be re-expressed as a the presence of both an electric field and magnetic field, Fermi surface property in terms of the Berry gauge con- with E · B 6= 0 [84]. Recall that that the Berry gauge nection: field on the Fermi surface is still defined in a general EFL. Z Here we will show that when this Berry gauge field has ΦB = dθAθ (56) non-trivial Chern number in a 3D EFL, then the system exhibits the chiral magnetic effect. Note that the integrand (as opposed to the integral) on We start from the anomaly equation that is the analog the right hand side – while only involving the Fermi sur- for an EFL in three spatial dimensions of Eq. (17) (which face – is not by itself invariant under k-space gauge trans- was stated for an EFL in two spatial dimensions); for formations. Going beyond non-interacting systems, for convenience we write it in an index-free form as: an interacting Fermi liquid it has been argued [83] that m d(∗j) = F ∧ F ∧ F, (58) there will be additional Fermi surface contributions to 48π3 16 where d is the exterior derivative on forms, “∗” is the invariance of the ground state immediately implies that Hodge star operator, and J is the current expressed as ν↑ = ν↓ := ν. The total charge density is then ρ = 2ν. a 1-form, and F is the 2-form gauge curvature (which Now, if we consider a spinful version of a Fermi liquid, we can write locally as F = dA, although A may not be the quasiparticle charge at each Fermi surface point will globally defined). Note that all these forms live on a 6- be conserved, as will the total quasiparticle spin. Note in dimensional space, M ×F, where M is the 4-dimensional particular that the quasiparticle spin at each Fermi sur- space-time manifold and F is the two-dimensional man- face point is not separately conserved due to the presence ifold that parameterizes the Fermi surface. of Landau interactions in the spin channel. Thus, for a Now suppose that we define the total charge current spinful Fermi liquid, the emergent symmetry group GIR jEM, which is a 1-form on M, by integrating j over the is a quotient of LU(1) × U(2), where the quotient corre- Fermi surface F. Suppose furthermore that we write F = sponds to identifying two U(1) subgroups: the U(1) sub- ∗ ∗ iθ πF (FEM) + πM (FBerry), where FEM are 2-forms on M group of U(2) corresponding to unitary matrices e I2, ∗ ∗ and F respectively, and πF and πM are the pull-back where I2 is the (2 × 2) identify matrix; and the U(1) R operators associated with the projections πF : F × M → subgroup of LU(1) generated by Nˆ = nˆ(θ)dθ, where F and πM : F × M respectively. In index notation, this nˆ(θ) are the generators of LU(1) as in previous sections. would just be saying that F is the sum of two terms, each Therefore, we define a spinful EFL to be a system which of which only depends on, and only has components in, has the same emergent symmetry group GIR. We as- M and F respectively. Then by integrating Eq. (58) we sume that the translation symmetry continues to embed find that into LU(1) according to Eq. (14), which defines the Fermi C surface of the EFL. Moreover, the mapping of the micro- d(∗j ) = F ∧ F , (59) EM 8π2 EM EM scopic U(2) into GIR is induced from a homomorphism ϕ : U(2) → U(2)IR, where the target U(2)IR is the corre- where sponding subgroup of GIR. One can show (for example, 1 Z by considering the induced homomorphism on the Lie al- C = F (60) 9 2π Berry gebras), that such maps are characterized by an odd F integer q, which again we interpret as the charge of a is the Chern number of the Berry curvature. We can also Fermi surface quantum (which in a Fermi liquid would write Eq. (59) as be a Landau quasiparticle). Then, we have, in particu- ˆ C lar that the microscopic charge Q corresponds in the IR ∂ jα = E · B, (61) theory to α EM 8π2 ˆ ˆ ˆ indicating that the charge is not conserved when the Q ∼ q(N↑ + N↓), (62) right-hand side is nonzero. where Nˆ↑ and Nˆ↓ are the generators of the U(1)↑,IR × U(2)↓,IR ≤ U(2)IR symmetry corresponding to the diag- onal matrices in U(2) . VII. EXTENSION TO SPINFUL SYSTEMS IR Now we can discuss the form that Luttinger’s theorem must take in a spinful EFL. We imagine inserting a 2π Throughout this paper, we have assumed that the only flux of U(1)↑,IR or U(1)↓,IR; this leads to a projective microscopic internal symmetry is U(1), the charge conser- representation of LU(1) described by Eq. (26) with some vation symmetry. Physically, one often wants to consider integer anomaly coefficient m↑ or m↓ respectively. U(2) systems which also have an SU(2) spin rotation symme- invariance again implies that m↑ = m↓ := m. Therefore, try. In that case, the full internal symmetry group is repeating the argument for Luttinger’s theorem, and tak- U(2), which acts by matrix multiplication on the vector ing into account Eq. (62), we find † † (ψ↑, ψ↓) of microscopic spin-up and spin-down electron mq creation operators. Here we will briefly discuss how the 2 VF = ν (mod 1). (63) general considerations above get extended in that case. (2π) (One is still allowed to use the results of previous sec- where VF is the volume enclosed by the Fermi surface. tions by simply ignoring the spin rotation symmetry, but With respect to the total charge density ρ = 2ν, on the taking into account the additional symmetry will lead to stronger constraints.) For simplicity we will focus on the case of d = 2 spatial dimensions. 9 The restriction to odd integers did not come up in Section VI, Firstly we observe that U(2) has a subgroup U(1)↑ × but we would have found the same restriction if we assumed Qˆ ˆ U(1)↓ ≤ U(2) corresponding to the diagonal unitary ma- that microscopically the fermion parity is (−1) where Q is the total charge (as is the case in an electronic system), and moreover trices. We write the generators of U(1)↑ and U(1)↓ as ˆ ˆ required (as is the case is all known examples) that an elementary Q↑ and Q↓; they measure the total number of up-spin Fermi surface quantum be fermionic. In the spinful case, the and down-spin electrons respectively. The total charge restriction to odd q comes directly from assumed form of GIR, is the sum Qˆ = Qˆ↓ + Qˆ↑. They both have correspond- although in principle one could get even q if we replaced U(2) with SO(3) × U(1) in the definition of G . ing fillings ν↑ and ν↓ in the ground state; however, U(2) IR 17 other hand, we have VIII. FERMI SURFACES COUPLED TO DYNAMICAL GAUGE FIELDS

2mq In this section, we will consider cases where the IR VF = ρ (mod 2). (64) (2π)2 theory comprises a Fermi surface coupled to a dynamical gauge field. As we will see, these do not constitute EFLs in the strict sense of Section VI, but they are still closely The extra factor of two takes into account the two pos- related and it still turns out to be helpful to think of the sible spin values, and agrees (setting m = 1) with the behavior of these systems in terms of their “kinematic” usual result for spinful Fermi liquids. (Thus, if we had properties. defined m with respect to the total charge as we did in We will consider three main examples. The first two Section VI, we would have found m = 2 for a spinful examples relate to a system of interacting electrons in two Fermi liquid/EFL). spatial dimensions with continuous translational symme- try in a magnetic field B, such that the magnetic filling Our discussion goes through with little modification ν = 2πρ/B (where ρ is the electron density) is equal to for a Kondo lattice in which itinerant electrons couple to ν = 1/2. Such a system of electrons is believed to form a a local spin-1/2 moment in each unit cell. In this case, “composite Fermi liquid” with a Fermi surface of “com- ˆtot we can define a U(1)↑ symmetry generated by Q↑ := posite fermions” which are distinct from the microscopic ˆz ˆ ˆz S + N/2 + Q↑, where S is the total spin component of electrons. There are two competing proposals for the IR the local moments, Qˆ↑ is defined as before with respect to theory of the system, which we consider separately: the the itinerant electrons, and N is the total number of unit one proposed by Halperin, Lee and Read (HLR) [70]; and cells. Assuming no spin ordering of the local moments, the one proposed by Son [85]. In the HLR and Son theo- z ˆtot we have that hS i = 0 and hence the total filling of Q↑ ries, the system is not quite a conventional Fermi liquid, tot even when expressed in terms of the composite fermions, is ν↑ = 1/2 + ν↑, where ν↑ is the contribution from the itinerant electrons. If we now apply similar arguments to because in the IR theory the composite fermions still cou- before, we find that Luttinger’s theorem for a putative IR ple to a dynamical gauge field. Fermi liquid requires a “large” Fermi surface that counts The final example relates to a in two the local moments as part of the Fermi sea, as previously spatial dimensions with discrete translation symmetry discussed with the flux threading argument in Ref. [7]. and no magnetic field. In certain circumstances, it is believed that the IR theory of such a system can have a For an alternative point of view, we can think of the “spinon Fermi surface” consisting of emergent fermions Kondo lattice model as being obtained from an Ander- coupled to a dynamical gauge field. son lattice model with a correlated half-filled band of electrons coupled to a separate weakly correlated par- tially filled band. Then, by definition, the charge ρ per A. HLR theory of composite Fermi liquid unit cell includes all the microscopic electrons, in par- ticular the contribution from the correlated band. If the We begin by discussing the microscopic symmetries of IR theory is a spinful Fermi liquid (or spinful EFL), then electrons in 2d in a uniform magnetic field. It is well Luttinger’s theorem in the form Eq. (64) shows that there known that in the presence of such a magnetic field, the Fermi surface must be “large” as in the previous para- translation operators Pˆx and Pˆy do not commute. Rather graph. Since our discussion (as well as that of Ref. 7) of they satisfy Luttinger’s theorem does not rely on perturbation theory ˆ ˆ ˆ in the interaction strength, it will hold even in the Kondo [Px, Py] = iBQ, (65) limit of the Anderson model. [Pˆx, Qˆ] = 0, (66)

Note that the arguments of this section assume that all [Pˆy, Qˆ] = 0, (67) the relevant symmetries – charge conservation, spin ro- tation and translation symmetry – are preserved and not where Qˆ is the operator for the total (gen- spontaneously broken, and that the IR theory is a spin- erator of the global U(1) symmetry). This symmetry al- ful Fermi liquid or at least a spinful EFL with the same gebra will have to be matched by any putative IR theory. emergent symmetries. If any of these conditions are vio- In an infinite system, we think of these operators Qˆ,Pˆx,Pˆy lated then the appropriate statement of Luttinger’s the- as the generators of the symmetry action on local observ- orem will be modified and in such cases, a “small” Fermi ables; hence, there is an ambiguity in their definition in surface could be permitted in the Kondo lattice model. that one is free to shift any of them by a constant with- The case of ‘Fractionalized Fermi Liquids’ which have a out changing the action on local observables. Thus, one small Fermi surface while preserving all the symmetries is free to fix them to have zero eigenvalue in the ground requires separate discussion but can be easily understood state. within the framework of this paper. We illustrate this for Let us now consider the HLR theory of the composite 1 simple class of such phases in Section X. Fermi liquid ground state at ν = 2 . Our starting point is 18 the Lagrangian for the HLR theory, which can be written expect the magnetic flux to induce a projective represen- as [70, 86] tation of LU(1) 10. Hence, the emergent symmetry group is LU(1),e where LU(1)e is obtained from LU(1) by replac- 2 1 0 L = L[χ, a] − b ∧ db + (A − a) ∧ db (68) ing the commutation relation [ˆn(θ), nˆ(θ )] = 0 obeyed by 4π 2π the generatorsn ˆ(θ) of LU(1) with: where a and b are dynamical U(1) gauge fields, and A 1 Z [ˆn(θ), nˆ(θ0)] = 2i δ0(θ − θ0) nˆ(θ)dθ, (75) is the background gauge field for the microscopic U(1) 2π symmetry. Here χ is a fermion field carrying unit gauge charge of a, and L[χ, a] is the Lagrangian describing a and with those elements Uf ∈ LU(1) corresponding to 1 Fermi surface of these fermions coupled to the gauge field functions f : S → U(1) with non-trivial winding number 1 R a. This is the corrected version (see e.g., Ref. [86]) of the Wf = 2π ∂θf(θ) excluded. [Because otherwise Eq. (75) ˆ −1 ˆ Lagrangian initially written by HLR [70], which suffered would imply that Uf NUf = (1+2Wf )N, which is math- from an improperly quantized Chern-Simons term. ematically inconsistent unless Wf = 0.] From Eqs. (69) Varying with respect to the temporal components a0 and (71), it is clear that the microscopic charge Qˆ em- and b0 gives the constraints beds into the IR symmetry group according to Qˆ ∼ Nˆ. Meanwhile, the continuous translation symmetry, with nˆχ − Bˆb/(2π) = 0, (69) generators Pˆx and Pˆy, should embed according to ˆ ˆ −2Bb + (B − Ba) = 0 (70) Z Pˆα ∼ kα(θ)ˆn(θ)dθ, (76) wheren ˆχ is the density of the χ fermions, and Bˆa and Bˆb are the magnetic fields of the a and b gauge fields re- which defines the Fermi surface k(θ). spectively. Meanwhile, varying with respect to A0 allows We will not attempt to extend the general discussion us to identify the microscopic charge densityρ ˆ as of topological terms as in Section V C to this more non- trivial symmetry group LU(1).e However, we can at least ρˆ = Bˆb/(2π). (71) argue for a version of Luttinger’s theorem for the com- posite Fermi liquid. Unlike in the case of no magnetic We define the excess charge according to field, it will turn out not to be necessary to invoke the ’t Hooft anomaly of the IR theory; the result actually Z follows directly from the structure of the emergent sym- Nˆ = (ˆn − n )d2x, (72) χ χ metry group. From the action of translations on the low-energy the- where nχ is the charge density in the ground state. We ory, we find that also define the effective magnetic flux that the χ fermions 2V experience (they only couple directly to a) according to [Pˆ , Pˆ ] ∼ F N,ˆ (77) x y 2π Z 2 where Φˆ = Bˆad x, (73) Z dky(θ) VF = kx(θ) dθ (78) Using 2nχ = B and the constraints above, it follows that dθ is the volume enclosed by the Fermi surface. We empha- size that here Nˆ is not just a number but an operator Φˆ 2Nˆ = − . (74) which acts non-trivially on the states of the IR theory. 2π On the other hand, microscopically we know that we Now we need to consider what form the IR symmetry must have the magnetic translation algebra of Eqn. 65 takes. Let us first consider what happens if we treat the with which we must impose compatibility of Eq. (77). gauge fields a and b at the mean-field level. In that case, This gives the χ fermions do not experience a magnetic field in the VF = πB, (79) ground state and therefore form a Fermi liquid for which the charge at each point of the Fermi surface is conserved. Thus, the emergent symmetry group is still something re- sembling LU(1) with generatorsn ˆ(θ). It is not exactly 10 One might ask why the magnetic flux does not also induce a LU(1), however, because if we consider low-energy exci- chiral flow of quasiparticles along the Fermi surface, violating ˆ R conservation ofn ˆ(θ). The point is that in that for a low-energy tations that carry a net charge N = nˆ(θ)dθ 6= 0, where excitation, the variation of the local charge density, and therefore nˆ(θ) are the generators of the emergent symmetry group, of the flux of a, should be spread over an arbitarily large volume, then from Eq. (74) we see that the χ fermions feel a net and hence the effective local magnetic field, which determines the magnetic flux. From the discussion of Section V D we velocity of the chiral flow, actually goes to zero. 19

which is Luttinger’s theorem for the composite Fermi liq- theorem. Note that, though it was suggested based on uid. numerics in Ref. [87] that Luttinger’s thoerem could be Now we should discuss to what extent we expect these violated in the composite Fermi liquid, more extensive considerations to survive once we include gauge fluctu- numerics showed this to be a finite size effect [88]. ations. The only property of the IR theory that we needed to derive Luttinger’s theorem was that the emer- gent symmetry group is LU(1).e The question is whether B. Son theory of composite Fermi liquid this emergent symmetry survives the inclusion of gauge fluctuations. Let us argue that, at any rate, the emergent We start from the Lagrangian [85, 86] symmetry group cannot be a compact Lie group, which 2 1 1 suggests that the full LU(1)e symmetry is preserved. We L = iχ¯Daχ− b∧db+ a∧db− A∧db+··· , (80) cannot apply the results of Section IV to show this, be- 4π 2π 2π cause if we define the magnetic unit cell such that the where a and b are dynamical U(1) gauge fields, and the discrete translations with respect to this unit cell com- first term represents the Lagrangian for a massless rela- mute, then the filling with respect to this unit cell is tivistic Dirac fermion field χ coupled to the gauge field a supposed to be ν = 1/2 for the composite Fermi liquid, 12. The “··· ” represents higher-order terms. This is the a rational number. corrected version [86] of the Lagrangian originally written However, we can instead argue as follows. Suppose by Son [85]. that the emergent symmetry group GIR is a compact Lie Proceeding similarly to Ref. [86], from this Lagrangian group. Then letp ˆx, pˆy, qˆ be the elements of the Lie alge- one obtains the constraints bra of GIR corresponding to the microscopic translation ˆ 1 ˆ 1 ˆ symmetries Pˆx and Pˆy and the microscopic charge conser- N − Φa + Φb = 0, (81) ˆ 4π 2π vation symmetry Q. Then since the microscopic symme- ˆ ˆ tries obey the algebra of Eqs. (65–67). it follows that the −2Φb + Φa = 0, (82) 1 same algebra must be satisfied byp ˆx, pˆy, qˆ. Eqs. (65–67) Qˆ = − Φˆ , (83) constitute the Heisenberg algebra and cannot be imple- 2π b mented inside of a compact Lie group unlessq ˆ is identi- ˆ ˆ ˆ cally zero. To see this, suppose that G is a compact where N is the charge of the χ fermions, Φa and Φb are IR the total magnetic fluxes of the a and b fields respectively, Lie group. Then it admits a finite-dimensional faithful ˆ ˆ ˆ ˆ ˆ unitary representation. Sincep ˆ , pˆ andq ˆ generate a Q is the total microscopic charge, and all of N, Q, Φa, Φb x y represent the excess values compared to the ground state. subgroup of GIR, this induces a finite-dimensional repre- ˆ ˆ ˆ sentation of the Heisenberg algebra Eqs. (65–67). Now From these equations we find that N = 0 and Φa = 2Φb. we invoke the famous fact that the Heisenberg algebra At the mean-field level the χ fermions form a Fermi does not admit any finite-dimensional representations in liquid, so the charge at each point on the Fermi surface whichq ˆ acts non-trivially11. is conserved. However, we have to impose the constraint that Nˆ = R nˆ(θ)dθ = 0. Assuming for the moment that Therefore, we conclude that, if G is a compact Lie IR the fermions feel no magnetic flux, this would imply that group then the charge conservation symmetry acts triv- the emergent symmetry group is ΩU(1) := LU(1)/U(1). ially on the IR theory; in other words, the system has Now if we take into account that the fermions feel an a charge gap and is an electrical insulator. This would effective magnetic flux given by Φˆ = 2Φˆ , we must have be inconsistent with the physics expected of a composite a b the commutation relation Fermi liquid. Thus, we expect that the emergent symme- try group remains LU(1)e in the presence of gauge fluctu- 1 Φˆ [ˆn(θ), nˆ(θ0)] = −2i δ0(θ − θ0) b . (84) ations. Of course, in principle it could go to some totally 2π 2π different group that is larger than any compact Lie group, but this does not seem very plausible, and moreover, as Since then ˆ(θ)’s are conserved, it must also be the case ˆ we we discussed in Section VI, the charge is still con- that the magnetic flux Φb is conserved and (since it is served at every point in the Fermi surface in the conven- quantized in units of 2π) generates a U(1)flux symme- tional description of the fixed point for a Fermi surface try. We can also observe that, for a general function coupled to dynamical gauge field. Thus, assuming that f : S1 → U(1), Eq. (84) implies that the corrspond- ˆ −1 ˆ ˆ the emergent symmetry group remains LU(1),e we find ing group element Uf satisfies Uf NUf = N + 2Wf Φb, that the composite Fermi liquid indeed obeys Luttinger’s

12 We follow the regularization convention adopted in Ref. [86], where the action for the massless Dirac fermion implicitly in- 11 To see this, just decompose into irreducible representations. In cludes the so-called “η invariant”. Physically this amounts to any such irreducible representation, by Schur’s Lemmaq ˆ acts assuming that the system also contains an additional massive proportionally to the identityq ˆ = QI for some scalar Q. Then Dirac fermion in addition to the massless one explicitly repre- taking the trace of [ˆpx, pˆy] = iBqˆ shows that Q = 0. sented in the Lagrangian. 20

1 R where Wf = 2π ∂θf(θ)dθ is the integer winding num- We see that the emergent symmetry group GIR takes the ber. Since we are imposing that Nˆ is identically zero form GIR = GIR,charge × SO(3), where the charge part (and Φˆb is not), we must exclude Uf with Wf 6= 0 from GIR,charge is identical to the emergent symmetry group the symmetry group, as in the HLR case. of the Son theory as described above [although for an electrical insulator the microscopic charge U(1) will act Hence, the full emergent symmetry group GIR is a cen- trivially on the IR theory instead of embedding into the tral extension of Ω0U(1) by U(1)flux where Ω0U(1) = emergent U(1) ], while the SO(3) part accounts for to L0U(1), and L0U(1)/U(1) is the subgroup of LU(1) corre- flux 13 sponding to functions f : S1 → U(1) with trivial winding the spin-rotation symmetry of the spinons . number. We note that this is a different group compared Unlike in the quantum Hall systems discussed above, to the HLR case, which reflects a physical difference be- for a Mott insulator one microscopically has only a dis- tween the Son and HLR theories. crete (commuting) translation symmetry. The discrete Eq. (83) tells us how the microscopic U(1) symmetry translations must embed into GIR in the usual way: embeds into GIR, and as before the continuous transla-  Z  tion should embed according to Eq. (76), which defines Tα ∼ exp −i kα(θ)ˆn(θ) , (86) the Fermi surface. By similar arguments to the HLR case one again finds Luttinger’s theorem Eq. (79) for the com- which defines the Fermi surface. From Eq. (85), the re- posite Fermi liquid. Since, again, the only property of the quirement that and commute give a non-trivial IR theory that we used was the nature of the emergent Tx Ty constraint: symmetry group, we expect that this result continues to hold in the presence of gauge fluctuations. V F = 0 or 1 (mod 1), (87) Note that our argument for Luttinger’s theorem in the (2π)2 2 Son theory is very different from the one given by Son [85], who noted that the density of composite fermions where VF is the volume enclosed by the Fermi surface. is given by ρCF = B/(4π), and then invoked Luttinger’s We can obtain more information by considering the ’t theorem for a Fermi liquid to relate this to the volume Hooft anomaly of the IR theory and invoking the filling of the Fermi surface. Of course, the composite Fermi constraints of Section III B. We assume that the system liquid is not a Fermi liquid beyond mean-field theory, but microscopically has spin-rotation symmetry, so that the the arguments of this paper have shown that Luttinger’s full microscopic internal symmetry group is U(2), and in theorem is in fact far more general. However, from this particular there is a U(1)↑ ×U(1)↓ subgroup as discussed point of view it remains unclear whether one ought to in Section VII, and we can define the corresponding fill- have expected a Luttinger’s theorem to hold with respect ings ν↑ and ν↓ which satisfy ν↑ = ν↓ := ν and ρ = 2ν, to ρCF , since the composite fermions do not carry charge where ρ is the filling of the total charge. For an insulat- under any global U(1) symmetry. The argument we gave ing state without symmetry fractionalization (see Section above does not suffer from these difficulties. X), the microscopic charge U(1) acts trivially on the IR theory and hence from the general theory of filling con- straints we find that ρ must be an integer. On the other C. Spinon Fermi surface hand, for a band insulator, ρ must be an even integer (taking into account the two spin components). Since a Finally, let us consider a system in 2 spatial dimensions Mott insulator by definition is an insulator that exists at where the IR theory consists of a spinful Fermi surface fillings where band insulators are impossible, we conclude coupled to a dynamical U(1) gauge field. This should ap- that ρ for a Mott insulator must be an odd integer and ply, for example, to Mott insulators exhibiting a “spinon hence ν = 1/2 (mod 1). Fermi surface”. From the general discussion of Section III B, we now see If we treat the dynamical gauge field at the mean-field that if we apply a flux Φ↑ of U(1)↑ and flux Φ↓ of U(1)↓, level, then the fermions form a spinful Fermi liquid de- it must be the case that translations act projectively on scribed by the considerations of Section VII. Thus, we it, have conserved Fermi surface densityn ˆ(θ). In particular, −1 −1 = (−1)(Φ↑+Φ↓)/(2π) (88) if we use the observation of that section that a spinful TxTyTx Ty Fermi surface has anomaly coefficient m = 2 with respect Now we need to see how this comes about in the IR the- to the total charge U(1), we conclude that then ˆ(θ)’s must ory. We return to Eq. (85) but take into account that, satisfy at the mean-field level where we can treat the dynamical gauge field as a background, the fermions now feels an 1 Φˆ [ˆn(θ), nˆ(θ0)] = −2i δ0(θ − θ0) b , (85) 2π 2π where Φˆ b is the operator that measures the magnetic flux 13 It is SO(3) rather than SU(2) or U(2) because of the constraint of the dynamical gauge field. Moreover, we have to im- that the total gauge charge, i.e. the net spinon number, is zero pose that the total gauge charge is zero, so R nˆ(θ)dθ = 0. for any state in the IR theory. 21

effective U(1)↑ flux of Φˆ b +Φ↑ and an effective U(1)↓ flux ARPES measurements observe “Fermi arcs” rather than 14 of Φˆ b + Φ↓. Hence, Eq. (85) gets generalized to a closed Fermi surface (see, eg, Ref. [89] for a review). One possible explanation is that these materials still have 1 2Φˆ + Φ + Φ a closed Fermi surface, but for some reason parts of the [ˆn(θ), nˆ(θ0)] = −i δ0(θ − θ0) b ↑ ↓ , (89) 2π 2π Fermi surface are not easily visible to the ARPES probe. On the other hand, 3-dimensional Weyl ex- which indeed gives Eq. (88) provided that hibit Fermi arcs at their 2-dimensional boundaries [90], but these can only exist at the boundary of a gapless 3- V 1 F = (mod 1), (90) dimensional bulk. It is an important fundamental ques- (2π)2 2 tion to ask whether a system with true Fermi arcs (not related to the above mechanisms) can ever exist. which is Luttinger’s theorem for a spinon Fermi surface Specifically, in this section we will ask whether it is As usual, we must consider to what extent we expect possible to extend the framework of EFLs to substitute these results to hold beyond mean-field theory. The re- a closed Fermi surface with a Fermi arc. We will find that sult Eq. (87) only depends on the structure of GIR, so we there is an obstruction to doing so. Therefore, in light must ask whether GIR remains the same upon including of the results of Section IV, the only remaining route to gauge fluctuations. Since we are assuming the system is Fermi arcs, assuming that they can exist in systems at a Mott insulator, with integer filling of the microscopic generic filling, would be to find some other way to have U(1) charge, we cannot use the results of Section IV. an emergent symmetry group larger than any compact Still, as we mentioned in the introduction to Section VI, Lie group, different from what occurs in EFLs. the charge on each patch of the Fermi surface remains Let us consider a theory in d spatial dimensions de- conserved in the usual approach to Fermi surfaces cou- scribed by a suitable generalization of the EFL class dis- pled to a dynamical gauge field, suggesting that GIR is cussed in Section VI. Specifically, we imagine that the indeed robust to gauge fluctuations and that Eq. (87) is Fermi surface is a parameterized by a (d−1)-dimensional satisfied generally. To get the stronger result Eq. (90), manifold F ; but, instead of requiring that F be a closed we also have to assume that the ’t Hooft anomaly of the manifold as before, we allow it to be a manifold with IR theory, captured by Eq. (89), remains unchanged by boundary. Then we assume that the emergent symme- gauge fluctuations. However, the classification of ’t Hooft try group is LF U(1), the space of smooth maps from anomalies is discrete – more concretely, there is no way to F to U(1). We can discuss ’t Hooft anomalies as be- continuously deform Eq. (89) without leading to incon- fore, and again we find that the ’t Hooft anomaly should sistencies – so it seems unlikely that the ’t Hooft anomaly be described by inflow from a Chern-Simons term on a would be affected by gauge fluctuations. Moreover from higher-dimensional manifold, for example for d = 2: Eq. (87) the only possibility other than Eq. (90) would be 2 m Z VF /(2π) = 0 (mod 1), in which case the Fermi surface is S[A] = 2 A ∧ dA ∧ dA (91) singular and presumably unstable. 24π M+×F

where M+ is an extension of the d+1-dimensional space- IX. IMPOSSIBILITY OF FERMI ARCS AND time manifold to a d + 2-dimensional manifold. Now, OTHER CONSTRAINTS ON THE FERMI however, we immediately encounter a problem if F has SURFACE a boundary ∂F , because the Chern-Simons term is not gauge-invariant – that is, the system has a ’t Hooft A. Fermi arcs anomaly – in the presence of boundaries (unless m = 0). At the boundary of M+, the gauge variation is canceled by the ’t Hooft anomaly of the EFL, but since the inte- In the celebrated pseudogap normal phase of the rior of M is supposed to correspond to a gapped SPT cuprate high temperature superconducting materials, + bulk, there is no way to cancel the anomaly at M+ × ∂F . Should we be worried about this anomaly? In gen- eral, an uncanceled ’t Hooft anomaly signals a violation

14 of charge conservation. We can see this concretely in We remark that if we set Φ↑ = Φ↓ = 2π, then the effect on ˆ ˆ ˆ spatial dimension d = 2 if we assume a particular model Eq. (89) can be absorbed by redefining Φb → Φb − 2π. Since Φb always has such an ambiguity in terms of its action on a back- of a Fermi arc which has Fermi liquid-like quasiparticles. ground flux (such ambiguities are the reason why fluxes are able Then in the presence of a magnetic field, the quasipar- to carry projective representations in the first place), we con- ticles have a chiral flow along the Fermi surface as dis- clude that Φ↑ = Φ↓ = 2π does not carry a non-trivial projective cussed in Section V D; but when they reach the end of the representation. This corresponds to the statement that the pro- Fermi arc they have no choice but to simply disappear. jective representations of G are classified, whereas IR,charge Z2 This violates conservation of microscopic charge (if the the projective representations of LU(1) (see Section V D) were Z classified. These results are consistent with the fact that the to- microscopic charge q of a quasiparticle is nonzero) and tal microscopic charge U(1) acts trivially on the IR theory since conservation of momentum (even when q = 0). Beyond it is an electrical insulator. this simple model, the ’t Hooft anomaly indicates the 22 same issue with lack of charge/momentum conservation, (a) (b) +1 and therefore we must impose that, if F has a boundary, +1 +1 then m = 0. However, by the discussion of Section V D this would imply the filling ν = 0 mod 1, so it cannot correspond to a state that exists at generic filling. Of course, this is only a proof of the impossibility of Fermi arcs if we assume that the system is described by the EFL framework. We cannot make more general state- +1 +1 ments, except to recall from Section IV that any system  +2  that can exist at generic filling yet is not an EFL needs to find some other way of having an emergent symmetry group that is larger than any compact Lie group. (c) (d) Finally, we should address one physical mechanism +1 +1 that one might think of to generate Fermi arcs, where the system couples to a critical boson at some wave-vector q. Then, in the quasiparticle picture described above, one +1 could imagine that a quasiparticle, upon reaching the end of a Fermi arc at momentum k, jumps to the beginning of another arc at momentum k−q while emitting a boson of momentum q. In the language of emergent symmetries,   this would correspond to introducing another operator Nˆboson that generates a U(1) group, and then writing the translation symmetry as (e) (f) +1  Z   ˆ  Tα = exp −i kα(θ)ˆn(θ)dθ exp −iqαNboson . (92) +1 This would amount to postulating a slightly different emergent symmetry group, namely a modification of LU(1) where one allows the corresponding functions f : +1 S1 → U(1) to have at least a finite number of discon-  +2  tinuities. The question is whether there is a meaning- ful non-trivial anomaly that one could write for such a group, as is necessary if it is to describe a system with FIG. 2. Allowed () and disallowed () Fermi surface config- urations in a two-dimensional Brillouin zone. Each Fermi sur- non-integer microscopic filling. It seems likely that there face segment has an associated anomaly coefficient m, shown is not, for reasons that we will explain at the end of the by the number next to the segment. An arbitary choice of ori- next subsection. entation of the segment (denoted with arrows) determines the convention for the sign of m. For allowed configurations, it is always possible to define an integer function n(k) – depicted B. More general constraints here by the different shading colors of different regions – such that the boundary between two regions carries the anomaly coefficient determined by the difference of n(k) between the Returning to EFLs, we can can also obtain some more two regions. In a Fermi gas, n(k) represents the number of general constraints than just the impossibility of Fermi occupied bands at the point k. arcs. Indeed, the most general statement will be that the Fermi surface must “enclose” a volume in a suitable generalized sense. This obviously fails for Fermi arcs, but it also rules out Fermi surfaces which are closed but account the orientation of the segment) intersecting at a nevertheless fail to enclose a volume in the Brillouin zone, given point must be zero. This is the most general sense e.g. see Figure 2(c). in which the Fermi surface must be “closed”. Thus, the Let us imagine generally that we construct the Fermi configuration shown in Figure 2(f) is allowed, but the one surface out of a collection of “patches” which might be shown in Figure 2(b) is not. glued together along their boundaries and might not all In fact, we can make an even stronger statement: con- have the same values of the anomaly coefficient m. Then figurations such as the one shown in Figure 2(c), even the requirement that the Chern-Simons term be gauge- though the Fermi surface is closed, are also disallowed invariant places a constraint on how patches are allowed (except possibly in the case where the Fermi surface to be glued together. For example, in two spatial di- quanta are uncharged, i.e. q = 0). In this case, the mensions the rule is that the sum of the anomaly coeffi- Chern-Simons term is gauge-invariant, but there is still cients for all the segments of Fermi surface (taking into an issue with the implementation of translation symme- 23 try. Recall that for d = 2 the translation symmetry Tx a filling ν by considering the anomaly of the emergent and microscopic charge operator Qˆ can be expressed in symmetry group. For Fermi surfaces that do not enclose terms of the LU(1) generatorsn ˆ(θ) as a volume, there is evidently no meaningful way to de- fine an associated ν. It is for this reason that we do not  Z  ∼ exp −i k (θ)ˆn(θ)dθ (93) expect there to be any non-trivial anomalies associated Tx x with the modified symmetry group described in the last Z paragraph of Section IX A, again because it would lead to ˆ Q ∼ q nˆ(θ)dθ. (94) Fermi surfaces that do not enclose a volume and therefore have no meaningful ν associated with them. In light of the commutation relations Eq. (19), when act- ing on a 2π flux we have X. THE ROLE OF FRACTIONALIZATION ˆ −1 ˆ TxQTx ∼ Q + mqWx, (95) where The analysis of Sections III and IV is not sufficient to describe systems which admit “fractionalized” excita- 1 Z dk (θ) W = x dθ (96) tions, i.e. localized excitations which cannot be created x 2π dθ locally. To see this, consider a specifc and familiar exam- ple, namely a system of bosons on a lattice in two spatial is a winding number of the Fermi surface on the Brillouin dimensions with microscopic filling ν = 1/2. A symme- torus. However, microscopically the translation opera- try preserving ground state that is allowed is one where tor should commute with the charge, so we must impose the IR theory is equivalent to a Z2 [91–93]. that mqWx = 0. Another way to see this (which general- Let us first recall the physical properties of this state izes more easily to higher dimensions) would be impose and how it is allowed at filling ν = 1/2 filling. The conservation of microscopic charge on the electric field key property is that the ground state is topologically response discussed in Section VI B. ordered and admits fractionalized excitations, which we From such considerations, we can obtain the most gen- can label as 1, e, m, f, where e is the Z2 gauge charge, eral constraint on the Fermi surface. We find that it must m is the Z2 gauge flux, and f is the composite of e and be the case that each point k in the Brillouin zone can m. Moreover, these excitations can exhibit “symmetry be assigned an integer n(k), such that the boundary be- fractionalization” of the microscopic symmetries, which tween two regions with different n(k) carries a Fermi sur- roughly means that they carry fractional quantum num- face with anomaly coefficient m given by the difference bers. Specifically, in the ν = 1/2 Mott insulator, one of ∆n. In a Fermi gas, n(k) has a natural interpretation: it the excitation types (say e) carries half-quantized charge describes the number of bands that are occupied at the under the microscopic U(1) symmetry. This is allowed point k. What we have found is that such an n(k) can because these excitations are fractionalized and can only always be defined in any consistent EFL. Note that in be created in pairs. Moreover, the m particle also expe- general the Fermi surface only determines n(k) up to ad- riences translational symmetry fractionalization[93–95], dition of a k-independent integer. Repeating the analysis −1 −1 which is to say that TxTyTx Ty = −1 when acting on of Section V D in this more general setting, we can also a single m excitation. Again, this does not contradict now formulate the most general version of Luttinger’s the fact that translations commute microscopically, be- theorem: cause the global number of m particles is always even. q Z Physically, one can think of the translational symmetry d2kn(k) = ν (mod 1), (97) (2π)d fractionalization as saying that the ground state has an e particle in each unit cell; the −1 phase factors then where the integral is over the whole Brillouin zone. comes from the −1 braiding phase for m going around We remark that that a compact mathematical way to e. But then since the e particle has 1/2 charge under state the above results is if we think of the Fermi surface, the microscopic U(1), these e particles contribute a fill- together with associated anomaly coefficients m of the ing ν = 1/2. It is also instructive to ask how this state is different patches, as defining a chain (in the homology compatible with the momentum balance constraints from d d theory sense) ω ∈ Cd−1(T , Z), where T represents the flux threading introduced by Oshikawa [7]. The point is Brillouin zone. Then the requirement that the Chern- that adiabatic threading of a 2π flux through one cy- Simons term be gauge invariant implies that this chain cle of a torus shifts the ground state from one topologi- is closed, i.e. it has trivial boundary, ∂ω = 0; whereas, cal sector to another which is at a different total crystal from the winding number arguments, we find that this momentum[5]. This change of the ground state momen- chain must be exact, ω = ∂κ for some κ. tum is then able to match what is expected from flux Finally, we note that in retrospect the result should threading. have been clear, because however the microscopic trans- How should we think about the emergent IR symmetry lations act in the IR theory (which is specified by speci- of this state? Deep in the IR, at scales below the gap to fying the Fermi surface), it must be possible to compute all quasiparticle excitations, we can describe it as a purely 24 topological theory. Naively the only emergent symmetry less theory Cgapless without any topologically non-trivial of the resulting theory might seem to be GIR = Z2, gen- excitations and a gapped topological theory Cgapped. In erated by the “electromagnetic duality” which exchanges that case, the microscopic filling ν is simply the sum of a Z2 gauge charge with a gauge flux; however, for the the contributions from each component: theory under consideration, neither translations nor the microscopic charge U(1) map into this duality symmetry. ν = νgapless + νgapped, (99) Therefore, following the analysis of Section III we would immediately conclude that ν = 0 mod 1. How then, are where νgapless is computed according to Section III and we to make sense, in the context of our general frame- νgapped is computed according to the previous paragraph. work, of the fact that such a IR theory actually can exist An example of such a case is the FL* Fermi liquid dis- at ν = 1/2? cussed in [8, 9]. We have to think about what went wrong in the argu- If we want to construct a more general theory beyond ment of Section III B. The assumption we made there is these examples, we might start to become a bit more that a 2π flux of the microscopic symmetry generated by uncertain about what exactly we mean in general by a the microscopic charge Qˆ will correspond in the IR theory “topologically non-trivial excitation”. Instead, we will introduce an alternative formalism that is more readily to a 2π flux of QˆIR, the generator of the corresponding IR symmetry. In any system with a charge gap, we will generalizable [96–99]. To motivate this formalism, note that in the Z2 gauge theory example, GIR = Z2 is not have that QˆIR is identically zero, so we would conclude that such an object would be completely trivial, lead- in fact the only emergent symmetry of the IR theory. Indeed, gapped topological phases in two spatial dimen- ing to a trivial filling. However, in the Z2 gauge theory example, a 2π flux of the microscopic symmetry is ev- sions, such as Z2 gauge theory, have emergent 1-form idently a non-trivial object since acting on it we have symmetries [25]. This is a general feature of gapped topo- −1 −1 logical phases. In fact one can think about the 1-form TxTyTx Ty = −1. In fact, as a defect of the IR the- ory it is equivalent to an m particle. In general, an in- symmetry operators, which act on closed strings when terpretation of non-trivial symmetry fractionalization is d = 2, as becoming the open string operators that cre- precisely the statement that certain flux configurations ate topologically non-trivial excitations at the endpoints of the microscopic symmetry look like topologically non- when the string is broken. trivial excitations of the IR theory. Therefore, it will Once one has introduced the emergent higher-form always be necessary to take into account symmetry frac- symmetries, one can phrase symmetry fractionalization tionalization when computing the filling. entirely in terms of these symmetries without needing to Let us first recall a few cases from previous works where talk about excitations at all. In general, a d-dimensional symmetry fractionalization of the microscopic symmetry system can have emergent 0-form, 1-form, ..., up to affects the filling. The first case generalizes the decon- (d−1)-form symmetries. These symmetries combine into fined gauge theory for bosons at ν = 1/2 (as discussed a structure called a “d-group” [98, 100–102], which we Z2 denote G . Meanwhile, the microscopic symmetries can above) to situations where the IR theory is an arbitrary IR also be thought of as a forming a d-group G , although gapped (2+1)-D gapped topological phase. The compu- UV tation of ν has been discussed in Refs. [10, 11]. The basic normally we will consider the case where the microscopic idea is that the translational symmetry fractionalization symmetries are all 0-form and there are no non-trivial higher-form symmetries. Then the correct way to de- is described by a class in H2(Z × Z,A) =∼ A, where A is the group of Abelian anyons. Roughly, we can imag- scribe the action of the microscopic symmetries on the ine that there is a background a particle sitting in each IR theory is through an n-group homomorphism unit cell. Meanwhile, the symmetry fractionalization of ρ : G → G . (100) U(1) is described by an Abelian anyon b, which we can UV IR think of as the anyonic charge carried by a 2π flux. Be- One might think if the microscopic symmetries are all cause of the background of a particles, if b particles have 0-form, then we do not need to worry about the higher- non-trivial braiding phase Sab with a particles, then they form part of G , but it turns out that even in this case IR carry a projective representation of translation symme- the homomorphism ρ contains not just the data of how try. Thus, we find that the filling is given by the microscopic symmetries map into emergent 0-form 1 symmetries, but also the data of the symmetry fraction- ν = log Sab (mod 1). (98) 2πi alization of the microscopic symmetries, which can be A nice way to interpret this equation is that since a interpreted as relating to the interplay between the mi- 2π flux has non-trivial braiding Sab with an a parti- croscopic symmetries and the emergent higher-form sym- cle, then the latter must carry fractional charge qa = metries. 1 2πi log Sab (mod 1) under the U(1) symmetry. Therefore, It is instructive to consider another case, in addition since we have an a particle in each unit cell, the average to gapped topological phases, where higher-form sym- charge per unit cell is ν = qa. metries modify the relation between microscopic filling Building on the previous case, we can imagine a sce- and the properties of the IR theory, namely superfluids, nario where the IR theory is equivalent to stacking a gap- which spontaneous break the charge U(1) symmetry. A 25 superfluid in d spatial dimensions has an emergent U(1) tinger’s theorem” in various settings. (d − 1)-form symmetry, since the winding number of the In certain cases [105–107], attempts have to been order parameter along any closed curve is conserved in made to formulate Luttinger’s theorem in a way that the IR theory. Moreover, a superfluid is a compressible could apply beyond perturbation theory in the interac- phase which can exist at any filling. One can show that tion strength by expressing it in terms of the volume the filling is encoded in the symmetry fractionalization of enclosed both by traditional Landau Fermi surfaces on the translation symmetry by the (d − 1)-form symmetry. which the electron Green’s function G(k, ω) has a pole Physically this is reflected in the fact that a vortex sees at ω = 0 and by “Luttinger surfaces” in which ReG(k, ω) an “effective magnetic field” given by the charge density changes sign via a zero of G(k, 0) instead of a pole. [103, 104]. However, it has been shown that no such result holds Finally, it is necessary for us to address the question in general [108, 109]. In our work, we have defined a of whether symmetry fractionalization can allow for the Fermi surface in a general system by its emergent sym- spirit of Theorem 1 to be bypassed; that is, is it possible metries, rather than by any particular properties of the for a system whose emergent symmetry is described by electron Green’s function. Our version of Luttinger’s the- a compact Lie d-group to exist at generic filling? The orem is therefore very different from the one disproven by superfluid example above shows that we have to restrict Refs. [108, 109]. to systems in which the microscopic U(1) and translation A potential violation of Luttinger’s theorem has been symmetries are not spontaneously broken. In that case, discussed in the context of holographic models [110–113]. we conjecture that a compact Lie d-group symmetry is in- For a system that is holographically dual to a gravita- compatible with generic filling. To motivate this, firstly, tional theory that includes a black hole, the charge hid- we expect that in order to have generic filling, at the least den behind the event horizon does not contribute to the the system needs to have an emergent continuous (d−1)- volume enclosed by the Fermi surface. It is not clear, form symmetry (e.g. a 1-form symmetry in d = 2 or a however, whether such a scenario can ever actually oc- 2-form symmetry in d = 3). The usual way to get emer- cur at zero temperature, since an event horizon would gent continuous higher form symmetries would be to have be associated with a nonzero entropy density, seemingly an emergent deconfined gauge field. However, in d = 3, in tension with the Third Law of Thermodynamics. If for example (we return to the more subtle d = 2 case somehow the event horizon could be stabilized at zero below), we know that a deconfined U(1) gauge theory temperature, however, one could envision it being asso- has two 1-form symmetries (electric and magnetic) but ciated with some exotic emergent symmetry group that is no 2-form symmetry. We can contrast this with gapped larger than any compact Lie group but different from that topological phases in d = 3, which typically have both a of a Fermi liquid, in which case a violation of Luttinger’s 1-form symmetry and 2-form symmetry. theorem would be compatible with the considerations of We also recall that a deconfined U(1) gauge theory in this paper. d = 2 spatial dimensions is dual to a superfluid, and that Next we discuss the numerical study of the t−J model the only way for such a theory to be stable (other than by in Ref. [114]. By an analytic continuation of a high- coupling to charged gapless fermions, which would explic- temperature series expansion, Ref. [114] computed the itly break the 1-form symmetry) is if the monopole events properties of the t−J model down to around temperature (in the superfluid language, charge creation events) are T ∼ 0.2J and found that, for several different definitions forbidden by one of the microscopic symmetries, in which of the “Fermi surface” at finite temperature, the volume case this symmetry will be spontaneously broken. inside the Fermi surface appears to be plateauing at a Therefore it appears, at least in these well-known value less than the value predicted by Luttinger’s theo- mechanisms for obtaining emergent higher-form symme- rem. However, our arguments in favor of Luttinger’s the- tries, that there is no way to obtain generic filling with- orem only applied at zero temperature. Since the ground out spontaneously breaking either the microscopic U(1) state of the t−J model is unknown and could well exhibit or translation symmetry. Whether there is some less fa- exotic physics such as high-temperature superconductiv- miliar theory in d spatial dimensions with an emergent ity, the significance of these finite-temperature observa- (d − 1)-form symmetry that can achieve this, we leave as tions remains unclear, especially given the ambiguity of an open question. the definition of the Fermi surface at finite temperature and the fact that the results of Ref. [114] also contain hints of Fermi arcs (see Section IX A). XI. DISCUSSION: RELATION TO Finally, we discuss Ref. [115], which studied an SU(N) “VIOLATIONS OF LUTTINGER’S THEOREM” generalization of the t − J model in a 1/N expansion where a Fermi liquid mean-field solution becomes exact In this work, we have argued that Luttinger’s theorem, in the limit as N → ∞. Ref. [115] found that the fluc- or a suitable generalization, in fact should hold in great tuations appear to cause a violation of Luttinger’s theo- generality, both for Fermi liquids and at least a certain rem to leading order in 1/N. According to our general class of non-Fermi liquids. Therefore, we must contend arguments, a violation of Luttinger’s theorem must cor- with past works that have reported a “violation of Lut- respond to a change in the emergent symmetry group 26

(such that the system is no longer an EFL). However, face quantum” excitations of an EFL do not need to have the mean-field solution is an EFL, in particular a Fermi the nature of quasiparticles, which are certainly expected liquid, and it seems unlikely that fluctuations could give to be absent in the strange metal.) Identification of the rise to a new emergent symmetry not already present strange metal as being in the EFL class, or even a sig- while remaining perturbative. Therefore, assuming there nificant generalization of it, would have further profound is no error in the calculations of Ref. [115], it seems that implications which we explore in a subsequent work [116]. we must interpret their results as signaling a breakdown of perturbation theory, such that the radius of conver- gence in 1/N is zero. In that case, there is no reason to ACKNOWLEDGMENTS believe that Luttinger’s theorem is violated at all. We thank Zhen Bi, Meng Cheng, Andrey Gromov, Shamit Kachru, Ethan Lake, John McGreevy, Max Metl- XII. CONCLUSION AND OUTLOOK itski, Chong Wang, and Liujun Zou for helpful discus- sions. D.V.E. was supported by the EPiQS Initiative In this work, we have made significant progress towards of the Gordon and Betty Moore Foundation, Grant No. an understanding of strongly correlated metals, going be- GBMF8684. T.S. is supported by a US Department of yond both Fermi liquid theory and perturbation theory Energy grant DE- SC0008739, and in part by a Simons in the interaction strength. We have highlighted the es- Investigator award from the Simons Foundation. This sential role of the emergent symmetry of the IR theory work was also partly supported by the Simons Collabo- and its ’t Hooft anomaly. ration on Ultra-Quantum Matter, which is a grant from Going forward, we believe that our work leads to a the Simons Foundation (651440, TS). new perspective on studies of non-Fermi liquids: from our point of view, a crucial question one should ask about any potential non-Fermi liquid state is what its emergent Appendix A: Filling does not correspond to the symmetry group is, and if it is supposed to represent a boundary of an SPT compressible state, in what way the emergent symmetry group and its ’t Hooft anomaly enable the system to sat- isfy the filling constraints developed in this paper for a Here we will explain why, contrary to what one might continuously tunable filling ν. (In particular, in this case initially assume, filling in d spatial dimensions does not the emergent symmetry group must be larger than any correspond to a boundary of a bulk SPT in d + 1 spa- compact Lie group). tial dimensions, when viewed in terms of the micro- scopic symmetries. The relevant symmetry group is What one might expect to find is that in fact, every G = d × U(1). One can show [117–119] that the SPT non-Fermi liquid is simply an ersatz Fermi liquid (EFL) IR Z classification for such a symmetry group in D spatial di- as defined in Section VI or a slight variant of it (for ex- mensions is given by ample including fractionalization as discussed in Section X). It is a very important open question to determine d! whether there might be a fundamentally different way to d × Y r realize compressible metallic states, with a totally differ- CD × CD−r , (A1) ent emergent symmetry group (which must still be larger r=1 than any compact Lie group). If such a possibility were to be realized, we would still expect that consideration of where Cm is the classification of U(1) SPTs in m spa- the emergent symmetry group and its ’t Hooft anomaly tial dimensions, C×k denotes the product of k copies of will still be a powerful way to determine properties of the d C, and is the binomial coefficient. Physically, C system, as we found for EFLs. r D In particular, it is quite energizing to imagine how our correspond to “strong” SPT phases, i.e. those protected results apply to the strange metal phase in cuprates [89], by U(1) alone, while the remaining terms correspond to assuming that it is possible for it to extend all the way “weak” SPT phases that can be understood by stacking down to a critical point at zero temperature, and assum- layers of lower-dimensional strong SPT phases. ing that it can exist in a translationally invariant system If filling in d spatial dimensions could somehow corre- without disorder. The filling at which the critical point spond to an SPT in D = d + 1 spatial dimensions, then occurs does not appear to be near any particular rational it would have to be an SPT that inherently relies on all filling. Therefore, it is likely that the filling is generic and d translation symmetries. Therefore, the only term in can be continuously tuned depending on the microscopic Eq. (A1) that could be relevant is the r = d term, C1. parameters, which as we have noted, leads to very strong The problem is that there are no 1-dimensional SPTs constraints on the IR theory. Although the strange metal protected by U(1) alone, in either bosonic or fermionic has extremely exotic properties, there does not appear to systems. Therefore, C1 = 0. So we conclude that there be any obvious reason why it could not still be an EFL are no SPTs in d + 1 spatial dimensions that could cor- (In particular, we emphasize again that the “Fermi sur- respond to filling in d dimensions. 27

Appendix B: Translation symmetry fractionalization Appendix C: Filling constraints in general on a 2π flux in 2-D systems at fractional filling dimension from the topological action of SPTs

In this section, we will give a more detailed argument Here we will discuss the sense in which the microscopic for why a 2π flux in a 2-D system with fractional fill- filling can be computed from the IR theory in general spa- ing should transform projectively under the translation tial dimension d, generalizing the discussion of Sections symmetry, as discussed in Section III B. III A and III B. First of all, we know that there is a one- First of all, let us recall that in the presence of a mag- to-one correspondence between the ’t Hooft anomaly of netic field one should introduce gauge-invariant transla- the IR theory and an SPT in d + 1 spatial dimensions. tion operators in the following way. For simplicity of Since the SPT phase is topological, at long wave-lengths notation, let us assume that the system is defined in a GIR SPT phase in d + 1 spatial dimensions is described continuous space (although with only a discrete trans- by a (Euclidean) topological action S[M,A] that depends lation symmetry); similar considerations can be made in on the d + 2-dimensional space-time manifold M and a a tight-binding model. We introduce the generator of gauge field A for the GIR symmetry. This action can be gauge-invariant continuous translations according to evaluated on any oriented space-time manifold M. (In Z some cases, there may be some additional structures re- Pe = P − A(x)ˆρ(x)d2x, (B1) quired on M, for example fermionic systems require a spin structure or spinc structure.) This action is purely where A(x) is the magnetic vector potential,ρ ˆ(x) is the imaginary (i.e. has no real part), does not depend on a local density operator, and P is the generator of trans- metric on M and, provided that GIR is a compact Lie lations when A = 0. These translation operators satisfy group, depends only on the SPT phase and is indepen- the commutation algebra dent of the microscopic details of the SPT ground state. Z In general, the action is only defined modulo 2πi since 2 [Peα, Peβ] = iεαβ B(x)ˆρ(x)d x, (B2) this does not affect the amplitude e−S. Let τ1, ··· , τd ∈ GIR be the elements of the IR sym- αβ where, where B(x) = ε ∂αAβ(x) is the magnetic field. metry group corresponding to microscopic translations. We then define the discrete translation operators Tx and We can define a generalization of the function α from Ty by the previous subsections in the following way. We set the (α) space-time manifold M to be the d+2-dimensional torus Tα = exp(−ia · Pe), (B3) T d+2. We require that over the first d of the d + 2 direc- where a(α) is the lattice translation vector in the α di- tions of the torus, the gauge field A is flat, with the fluxes rection. through the associated 1-cycles given by τ1, ··· , τd. Then Now consider any state |ψi that obeys the “cluster over the final two directions on the torus, we require that property” there be a 2π flux of the U(1) subgroup of GIR (generated ˆ hρˆ(x)ˆρ(x0)i − hρˆ(x)ihρˆ(x0)i ≤ C|x − x0|−η (B4) by QIR) corresponding to the microscopic U(1) symme- try. Finally, we define α(QˆIR|τ1, ··· , τd) = −iS[M,A]. for some constants C and η > 0, and h·i = hψ| · |ψi One can show that this agrees with the previous defini- denotes expectation values with respect to |ψi. Then it tions of α in the cases d = 1, 2. R 2 follows from Eq. (B2) that, if we fix that B(x)d x = 2π Thus, our hypothesis for the form of the filling con- but make the magnetic field more and more spread out, straint is then the variance 1 2 2 ν = α(Qˆ |τ , ··· , τ ) (mod 1). (C1) h[Peα, Peβ] i − h[Peα, Peβ]i (B5) 2π IR 1 d goes to zero in the limit where the magnetic field is in- We have not devised a physical derivation analogous to finitely spread out. Therefore, when acting on such a the d = 1 and d = 2 cases; however we expect that this state |ψi, and in this limit, we can replace the commuta- should be possible, since α should relate to the anomalous tor by its expectation value: action of translations on a 2π flux, which in d spatial Z 2 2πν dimensions is a (d − 2)-dimensional object. [Peα, Peβ] ∼ h[Peα, Peβ]i = iαβ B(x)ρ(x)d x ∼ iαβ, v We can extend this definition to include higher form (B6) symmetries as follows. As we have discussed in Section where v is the volume of one unit cell and ρ(x) = hρˆ(x)i. X, in this case we can include the possibility of symmetry To get to the last identification, we used the fact that fractionalization by considering the map ρ : GUV → GIR B(x) is very slowly varying, and that the integral of ρ(x) to be a map of higher groups, where GIR contains all over one unit cell is ν. Finally, from Eq. (B6) and the higher form symmetries. We will show that there is a GIR definition Eq. (B3), we find that gauge field A on T d+2 associated to ρ which generalizes d+2 −1 −1 2πiν the A above, so that ν = −iS[T ,A]/2π. First, we TxTyTx Ty ∼ e , (B7) d+2 d equip the d + 2 torus T with a Z × U(1) gauge field which is Eq. (8). A0 defined as follows. For j = 1, . . . , d, the jth 1-cycle 28

d+2 R 0 Cj of T has A equal to the generator of the jth which amounts to coupling the zero Hamiltonian to the Cj factor. On the last two coordinates, which form a T 2, gauge field At = f(θ). Then we find that Heisenberg Z itH −itH we place a 2π flux in the U(1) factor. We then take evolutionn ˆ(θ, t) = e nˆ(θ)e satisfies 0 A = ρ(A ). d Z Note that, as we mentioned, in certain cases the topo- nˆ(θ, t) = i c(θ0, θ)f(θ0)dθ0. (D3) dt logical action could depend on other data (such as a spin structure for fermionic systems). However, we expect We should compare this with the anomaly equation that if the action still depends non-trivially on this ad- Eq. (D1), in which we choose Aθ to be independent of ditional data for the particular space-time manifold M time, we have ejθ = δH = 0, and we identify the time- and gauge field A considered here, this corresponds to δAθ component jt with the expectation value ofn ˆ(θ, t). We cases where the microscopic symmetry G = d × U(1) e UV Z find that, to be consistent with Eq. (D3), it must be that still has a non-trivial ’t Hooft anomaly in d spatial di- mensions. In this case, the system cannot be realized as lattice system in d spatial dimensions with translation Z mφ d i c(θ0, θ)f(θ0)dθ0 = − f(θ), (D4) symmetry and on-site U(1) symmetry, only as a bound- 2π dθ ary of an SPT in d + 1 spatial dimensions. In such sys- tems, the filling of the d-dimensional system does not which is true for an arbitrary test function f only if need to be well-defined. mφ c(θ, θ0) = −i δ0(θ − θ0). (D5) 2π Appendix D: Projective representation of loop Finally, to obtain the extra factor of q in the right-hand group on monopole from ’t Hooft anomaly side of Eq. (19), we recall that Eq. (13) implies that φ, which we can interpret as the number of flux quanta of R In this Appendix, we will show that Eq. (19) follows the IR U(1) symmetry generated by nˆ(θ)dθ, is q times from the anomaly equation Eq. (17). the number of flux quanta of the microscopic U(1) sym- Let us work on a 4-dimensional space parameterized metry. by coordinates (t, x, y, θ). We start by considering a con- figuration where the Ax and Ay components of the gauge Appendix E: Projective representation of loop group field depend only on x and y, and the At, Aθ components depend only on z,t. Then, we integrate Eq. (17) over x in a Fermi liquid and y. We obtain In this Appendix, we derive Eq. (19) in a Fermi liq- µ mφ uid by working in terms of the semiclassical theory of ∂µej = (∂tAθ − ∂θAt), (D1) 2π band electrons. Strictly speaking, we will only treat non- interacting electrons, but since our arguments are purely where the index µ now varies only over t and θ, and geometric and the result ultimately can be reduced to we have defined φ = 1 R (∂ A − ∂ A )dxdy, which is 2π x y y x a contribution coming solely from the Fermi surface, we the number of flux quanta passing through the system, µ R µ expect that the same results will hold in a Fermi liquid and ej = j dxdy. Effectively, what we have done is in light of the quasiparticle picture. compactify our original system in 2 spatial dimensions For clarity, in this appendix we temporarily work in to one in 0 spatial dimensions. This 0-D system has a ’t units where 6= 1. Then we can write Eq. (19) as Hooft anomaly that depends on φ, and is canceled by in- ~ flow from a 2-D SPT. Recall that such ’t Hooft anomalies Φmq [ˆn(θ), nˆ(θ0)] = −i δ0(θ − θ0) (E1) correspond to projective representations of the symmetry (2π)2~ group. If we set φ = 1, this corresponds to the projec- tive representation of the symmetry group acting on a 2π where Φ = R B(x)d2x is the total magnetic flux. There- flux. fore, in the semiclassical theory what we want to show is In our case, where the symmetry group is G = LU(1), the Poisson bracket we can express the projective representation through this Φmq central extension of the commutation algebra of the sym- {n(θ), n(θ0)} = − δ0(θ − θ0). (E2) (2π )2 metry generatorsn ˆ(θ), i.e. ~

0 0 Note that for each θ, n(θ) is a semiclassical observ- [ˆn(θ), nˆ(θ )] = c(θ, θ ), (D2) able, i.e. a real-valued function on the semiclassical phase space. where c(θ, θ0) is a c-number (i.e. not an operator) to be What we will actually prove is the following equivalent determined. statement: The projective representation should be independent of the choice of Hamiltonian, so we are free to choose Φqm Z dg(θ) R {nf , ng} = − f(θ) dθ, (E3) H = f(θ)ˆn(θ)dθ for some arbitrary test function f(θ), (2π~)2 dθ 29

for any smooth functions f and g, where we have defined where in this equation both the left- and right-hand sides have an implicit dependence on the phase space vari- (1) (N) Z ables (k , ··· , k , x(1), ··· , x(N)), on the right-hand nf = f(θ)n(θ)dθ. (E4) side through n(x, k). Now let us write n(x, k) = n0(k) + δn(x, k), where n0(k) has no implicit dependence on the phase space We can relate nf and ng to the density n(k, x) of semi- classical particles according to: variables and represents the equilibrium distribution of particles in the ground state (keeping in mind the Pauli Z 2 2 exclusion principle). For low-energy configurations, the nf = f+(k)n(k, x)d xd k, (E5) contribution of δn to Eq. (E11) will be negligible com- pared to n0. Hence, we can replace n with n0 in Eq. (E11) where f+(k) is any smooth extension of f away from the and we find Fermi surface such that f+(kF (θ)) = f(θ), where kF (θ) Z 2 is the momentum of the Fermi surface as a function of θ. {nf , ng} = −qΦ n0(k)[∇kf+(k)] × [∇kg+(k)]d k, Note that by n(k, x) we mean that for each k, x, n(k, x) (E12) is a function of phase space variables. If there are N par- Z ticles in total then the phase space variables are labeled 2 = −qΦ n0(k)∇k × [f+(k)∇kg+(k)]d k. by (x(1), ··· , x(N), k(1), ··· , k(N)), and we can write (E13) n(k, x)(x(1), ··· , x(N), k(1), ··· , k(N)) 2 Now we observe that n0(k) = 1/(2π~) in the occupied N region of the Brillouin zone, and 0 outside. So by Stokes’ X = δ2(k − k(i))δ2(x − x(i)). (E6) theorem we obtain i=1 qmΦ I {n , n } = − [f (k)∇ g (k)] · dk f g (2π )2 + k + Here k(j) denotes the gauge-invariant momentum. That ~ Fermi surface (E14) means that the Poisson brackets of the x(j)’s and k(j)’s qmΦ Z dg(θ) get modified due to the magnetic field, as we discuss be- = − f(θ) dθ, (E15) low. (2π~)2 dθ Now we can compute {nf , ng}. We will first do it for where m = ±1 (relative to an arbitrary choice of orien- (1) (1) a single-particle phase space (k , x ). Then we have tation of the Fermi surface, m = +1 or −1 corresponds to which side of the Fermi surface the occupied region of (1) (1) (1) nf (k , x ) = f+(k ) (E7) the Brillouin zone is on). This is the result we wanted to show. and hence

∂f (k(1)) ∂g (k(1)) {n , n } = + + {k(1), k(1)} (E8) Appendix F: Proof of Theorem 1 f g (1) (1) α β ∂kα ∂kβ (1) (1) (1) In this appendix, we give a proof of Theorem 1 from = −B(x )[∇k(1) f+(k )] × [∇k(1) g+(k )], (E9) Section IV. The result in fact follows directly from the fol- lowing mathematical property of the function α defined where we used the fact that the momenta k(1) obey the in Section III A for spatial dimension d = 1, in Section magnetic algebra 15 III B for d = 2, and in Appendix C for general d. Lemma 1. Let A be an Abelian subgroup of GIR. The (1) (1) (1) {kα , kβ } = −qB(x )εαβ (E10) function α(Q|τ1, ··· , τd) defined in Section III is linear in each argument when restricted to A. That is, for (here we have chosen the convention that k carries units 0 a1, ··· , ad, a1 ∈ A, and where Q is the infinitesimal gen- of momentum, not inverse length.) For general N, we erator of a U(1) subgroup of A, then we have have 0 Z α(Q|a1a1, a2, ··· , ad) = α(Q|a1, a2, ··· , ad) 2 2 0 {nf , ng} = −q B(x)n(x, k)[∇kf+(k)]×[∇kg+(k)]d xd k, + α(Q|a1, a2, ··· , ad) (mod 2π), (F1) (E11) and similarly for each of the arguments after the “|”. Proof. It is easy to verify this given the concrete defini- tions of α in the cases d = 1 and d = 2. In general d, 15 In general there there can be a Berry curvature correction to where α is defined as discussed in Appendix C, the result this equation [120], but it enters only at next-leading order in follows from the hypothesis that the topological action is the magnetic field strength, so we disregard it here. a cobordism invariant [50, 52, 56]. 30

Lemma 1 is all we need to give: Appendix G: Bounds on Filling Denominators for Compact GIR

Proof of Theorem 1. Let τ1, . . . , τd ∈ GIR be the IR symmetries corresponding to the microscopic translation Theorem 1 says that a compact GIR is inconsistent with an irrational filling fraction. In this appendix, we symmetries and let QˆIR be the infinitesimal generator of the IR symmetry corresponding to the microscopic U(1) will discuss how, given an assumed GIR, in a fixed space dimension d, one can compute a denominator N such rotation symmetry. Since the microscopic translations GIR commute with each other and with charge, τ , ··· , τ and that GIR is consistent only with rational filling fractions 1 d of the form m/N , for m ∈ . We will show the fol- Qˆ must all commute. Let A be a maximal Abelian sub- GIR Z IR lowing group of GIR (that is, a subgroup that cannot be enlarged as a subgroup of GIR while remaining Abelian) that con- Theorem 2. In d > 1 space dimensions, for any com- ˆ tains τ1, ··· , τd and the subgroup generated by QIR. One pact emergent symmetry GIR (which could be a subgroup can show16 that A is a compact Abelian Lie subgroup of of the full group of emergent symmetries) containing the GIR. Let CA be the number of connected components of translation generators τ1, . . . , τd and the subgroup U(1) A, and let A0 ≤ A be the connected component of the corresponding to microscopic particle-number conserva- CA identity element. Then since [a] = 1 for any element tion, the maximum denominator NGIR divides the largest CA [a] ∈ A/A0, we conclude that a ∈ A0 for any a ∈ A. order of elements in the torsion part of the (co)bordism tors d+2 tors Now since A is a compact Abelian Lie group, the group Ωd+2(GIR) = Ω (GIR) , ie. the group of r connected component A0 must be isomorphic to U(1) d + 1-space-dimensional SPT phases such that for some for some r. Hence, we can write elements of A0 as n, a stack of n copies of the SPT is trivial. (θ1, ··· , θr) for angular variables θ1, ··· , θr. We now re- ˆ Theorem 3. In d = 2 space dimensions, assum- turn to the function α(QIR|τ1, ··· , τd) that computes the ing a compact emergent symmetry G satisfying the ˆ IR filling. Henceforth, we will pick a given QIR and hold conditions of Theorem 2 and furthermore of the form it fixed, therefore we write simply α(τ1, . . . , τd). The GIR = H × U(1), the maximum denominator NGIR linearity of α established by Lemma 1 implies that α tors divides |π0(GIR)| · |π1(GIR) |, where π0(GIR) is the is continuous in each argument, and, moreover, that if tors set of connected components of GIR and π1(GIR) is a1 = (θ1, ··· , θr) ∈ A0, then the torsion part of its fundamental group. If there is

also a finite 1-form symmetry K, then NGIR divides r tors X |π0(GIR)| · |π1(GIR) | · |K|. α(a1, ··· , ad) = θjmj(a2, ··· , ad) (mod 2π), (F2) j=1 Theorem 4. In d = 3 space dimensions, assum- ing a compact emergent symmetry GIR satisfying the ×(d−1) for some continuous integer function mj : A → Z, conditions of Theorem 2 and furthermore of the form G = H × U(1), the maximum denominator N where we have suppressed the dependences on QˆIR, which IR GIR divides |π (G )| · gcd(|π (G )tors|, |π (G )|), where we hold fixed. But now continuity implies that mj is 0 IR 1 IR 0 IR π (G ) is the set of connected components of G and zero if any of its arguments are in A0. We therefore 0 IR IR π (G )tors is the torsion part of its fundamental group conclude that α(a1, ··· , ad) = 0 (mod 2π) if any two of 1 IR and gcd denotes the greatest common divisor. In partic- its arguments are in A0. Therefore: ular, for connected GIR, NGIR = 1, so a compact con- nected emergent symmetry is incompatible with any frac- (C )2α(τ , ··· , τ ) = α(τ CA , τ CA , τ , ··· , τ ) = 0 A 1 d 1 2 3 d tional filling. With a finite 1-form symmetry K, then (mod 2π), (F3) tors NGIR divides |π0(GIR)| · gcd(|π1(GIR) |, |π0(GIR)|) · gcd(|K|, π0(GIR)). If there is furthermore a finite 2 and hence the filling ν is an integer multiple of 1/(CA) . 2-form symmetry J, then NGIR divides |π0(GIR)| · tors Finally, to see that we can take NGIR finite for any gcd(|π1(GIR) |, |π0(GIR)|) · gcd(|K|, |π0(GIR)|) · |J|. given compact Lie group G , we note17 that for any IR We note that since Ω (G )tors is finite, Theorem 1 such G , there are only finitely many conjugacy classes d+2 IR IR follows from Theorem 2. of maximal Abelian Lie subgroups A ≤ G . IR Proof of Theorem 2: As argued in Appendix C, the filling fraction associated to GIR may be computed from For an alternative perspective that gives more precise the emergent GIR ’t Hooft anomaly as well as the data

bounds on NGIR , see the next appendix, Appendix G. of how U(1) and translation symmetries are realized in GIR. In particular, the translation generators τ1, . . . , τd and the U(1) generator Qˆ define a GIR gauge field on d+2 d 2 T = T × T , where the τj define the fluxes through d 16 By using the Closed Subgroup Theorem for Lie groups [121]. the d coordinate 1-cycles of T and Qˆ defines a 2π flux of 17 This follows [122] from the “neighboring subgroup theorem” of T 2. The emergent ’t Hooft anomaly is associated with a Ref. [123] cobordism invariant of GIR gauge fields on d+2 manifolds 31

(the partition function of the ((d + 1) + 1)D-dimensional Simons invariant is proportional to SPT associated with the ’t Hooft anomaly), and that Z d+3 cobordism invariant evaluated on this particular back- (dA) 2 . 2πiν ground equals e . D2×T d−1×T 2 Let us denote the group of cobordism invari- However, since the 2-form dA has only nonzero compo- ants by Ωd+2(G ). In the notation of [52], it is IR nents in 4 directions, D2 and T 2, its power in the inte- Ωd+2(B(G / F ), ξ), where ξ is an orientable bundle spin IR Z2 grand above vanishes if d + 3 > 4, ie. if d > 1 18. This whose 2nd Stiefel-Whitney class classifies the extension proves the claim. of G / F by fermion parity F . The group Ωd+2(G ) IR Z2 Z2 IR Proof of Theorem 3: We can get a handle on the max- sits in a (split) short exact sequence [56] imum denominator NGIR , which we have proven divides tors |Ωd+2(GIR) |, using the Atiyah-Hirzebruch spectral se- d+2 Hom(Ωd+3(GIR), Z) Ω (GIR) quence for computing the latter. This gives us an upper bound

tors d+2 Hom(Ωd+2(GIR) ,U(1)) Y b spin tors NGIR divides |Hj(BGIR, Ωd+2−j) |, j=0 where Ωd+2(GIR) and Ωd+3(GIR) are the associated bor- tors dism groups of manifolds and Ωd+2(GIR) indicates b F spin where GIR = GIR/Z2 and Ωk is the usual spin bordism the torsion subgroup of Ωd+2(GIR), elements x for which group, which in the low degrees of physical interest is there exists an n such that n · x = 0. spin The kernel of this extension, Hom(Ωd+3(GIR), ), is Z Ω0 = Z characterized by generalized Chern-Simons invariants, spin described by Chern-Weil forms evaluated on a bounding Ω1 = Z2 d + 3-manifold M with ∂M = T d+2. We will show that spin Ω2 = Z2 for d > 1 these invariants do not contribute to the fill- spin d+2 Ω = 0 ing fraction. More precisely, the map Ω (GIR) → U(1) 3 given by evaluating the cobordism invariant on such a spin Ω4 = Z d+2 T background as described above sends the kernel to spin Ω5 = 0. zero. Therefore, the maximum denominator NGIR divides tors tors |Hom(Ωd+2(GIR) ,U(1))| = |Ωd+2(GIR) |, which is This upper bound can be further improved by taking finite. This also gives an alternative proof of Theorem into account the known differentials of the dual spectral 1. In fact, NGIR divides the largest order of elements d+2 tors sequence for cobordism Ω (GIR), which has isomorphic in Ωd+2(GIR) , which we will use below. This group d+2 tors tors d+2 tors torsion, Ω (GIR) = Ωd+2(GIR) . See [124] for a is isomorphic to its cobordism partner Ω (GIR) be- review. cause the short exact sequence above splits. Another thing we can do is consider the special case Now we will show that for d > 1 the generalized Chern- G = H × U(1), where U(1) and H is compact. Be- Simons invariants do not contribute to the filling fraction. IR cause τ , . . . , τ and Qˆ must commute, they are always For even d, there are no generalized Chern-Simons invari- 1 d contained in such a group. In this case, by taking consid- ants and we are done, so let us consider odd d instead. ering d+2-manifolds of the form M ×T 2 with a arbitrary We will use a flat M so that there are no metric con- G bundle on M and a U(1) bundle with Chern num- tributions to the Chern-Simons invariant. We will also IR ber 1 on T 2, our d + 2-dimensional cobordism invariant use the fact that the remaining Chern-Simons invariants defines a d-dimensional cobordism invariant for M. By are defined for G bundles by taking trace in a unitary IR similar arguments as above, this gives us the further con- representation R : G → U(n) and is therefore equal IR straint that N divides the element of highest order in to a Chern-Simons invariant U(n) bundle associated to GIR Ω (G )tors = Ωd(G )tors. it by the map R. Then, we use the fact that τ , . . . , τ d IR IR 1 d For d = 2, using the Atiyah-Hirzebruch spectral se- and the U(1) generator Qˆ commute in G , hence their IR quence for Ω2(G ), we find in the case G = H × U(1), images in U(n) also commute and can be simultaneously IR IR N divides diagonalized. This reduces the calculation to an Abelian GIR n Chern-Simons invariant for the subgroup U(1) of diag- |Ωspin| · |H1(BGb , Ωspin)| · |H3(BGb , Ωspin)tors|. onal matrices in U(n). Taking the trace reduces this to 2 IR 1 IR 0 a calculation for a U(1) bundle. One can show that since the fermion parity is a sub- We can extend our U(1) bundle from T d+2 = T d × T 2 group of the U(1) factor, that the first two factors do to D2 × T d−1 × T 2, where D2 is a disc whose boundary is the first coordinate cycle of T d. The extended gauge field A is defined so that it has curvature only along D2 18 (integrating to TrR(τ1) as required by the extension) and We note that for d = 1, the Chern-Simons invariant is an integer along T 2 (where it has some Chern number). The Chern- multiple of TrR(τ1). 32 not contribute to the spin bordism because the spectral that we have some functional F[f], where f is a function sequence has a nonzero differential there. The third part of a single variable into R/Z that can be expressed as tors is |π0(GIR)| · |π1(GIR) | torsion, so f(x) = tg(x), (H1) N divides |π (G )| · |π (G )tors|. GIR 0 IR 1 IR and we we wish to extract the asymptotic dependence of We note we can often get even better bounds if we better F[f] as t → ∞ while keeping the function g fixed. First we imagine approximating f by its values on the understand the cobordism invariants for GIR. For exam- r discrete points x1, ··· , xk. Then we can expand F as a ple, if GIR = Z2 × U(1), one can show the maximum r Fourier series denominator is NGIR = 2, even though this group has 2 components. X 2πit(n g(x )+···+n g(x )) F[f] = a e 1 1 k k , (H2) To include a finite 1-form symmetry K, there is one n k more term which contributes to the spectral sequence, n∈Z 3 2 H (B K, Z) = K, which is |K| torsion. where the sum is over integer vectors n ∈ Zk. We can Proof of Theorem 4: For d = 3, GIR = H × U(1), the rewrite this as contributions to the Atiyah-Hirzebruch spectral sequence ∞ 3 X X for Ω (GIR) are F[f] = S(0) + S(1) + S(2) + ··· , (H3) n n1,n2 1 b spin 2 b spin 4 b spin tors n6=0 n16=0,n26=0 |H (BGIR, Ω2 )|·|H (BGIR, Ω1 )|·|H (BGIR, Ω0 ) |. where S(0) = a and As in d = 2, the first factor does not contribute. The sec- 0 k ond factor has one piece that can contribute, coming from X 2 2 b S(1) = a e2πitmg(xj ) (H4) the map H (BU(1), Z) → H (BGIR, Z2) but it turns out n (j:n) to give a spin Chern-Simons term for the U(1) factor and j=1 X we can disregard it, ie. it does not contribute to the tor- S(2) = a e2πit(n1g(x1)+itn2g(x2)) 3 n1,n2 (j1:n1,j2:n2) sion in Ω (GIR). The third piece can be bounded using j16=j2 the Serre spectral sequence for the extension (H5) b,0 b b S(3) = ··· , (H6) GIR → GIR → π0(GIR) = π0(GIR), n1,n2,n3 from which we find it is |π0(GIR)| · where (j1 : m1, j2 : m2), for example, is the vector n tors gcd(|π1(GIR) |, |π0(GIR)|) torsion, so obtained by setting nj1 = m1 and nj2 = m2 and nj = 0 for j∈ / {m1, m2}. By taking the limit as k → ∞ and the tors NGIR divides |π0(GIR)| · gcd(|π1(GIR) |, |π0(GIR)|). xj’s become dense, we find that S` → I`, where

Note that when GIR is connected, we re-establish the re- Z sult N = 1: connected, compact G are incompatible (1) 2πitng(x) GIR IR In = dx an(x)e (H7) with fractional filling in d = 3. When there is also a finite 1-form symme- Z I(2) = dx dx a (x , x )e2πit(n1g(x1)+n2g(x2)), 4 b spin n1,n2 1 2 n1,n2 1 2 try K, it contributes a term to H (BGIR, Ω0 ) 1 b 3 2 via H (Bπ0(GIR),H (B K, Z)), which is (H8) gcd(|π (Gb )|, |K|) torsion. 0 IR I(3) = ··· (H9) When there is also a finite 2-form symmetry J, it con- n1,n2,n3 4 b spin 4 2 tributes a term to H (BGIR, Ω0 ) via H (B J, Z) = Now we invoke the theory of stationary phase integrals J. which tells us that, for a function h(x) of a q-dimensional Finally we note that the bounds so derived hold for variable x, then as t → ∞ every subgroup GIR of the full emergent symmetry, so Z long as it contains the translation generators τ , . . . , τ q 2πith(x) X −q/2 2πitf(x∗) 1 d d xϕ(x)e ∼ cx t e , (H10) and the U(1) particle-number symmetry. Thus, one does ∗ x∗∈Σ not need to know the entire emergent symmetry to obtain a bound. Further, by varying GIR, one can sometimes for some constants cx∗ and where the sum is over the set find a better bound for the problem at hand. Σ of solutions to ∇f(x) = 0. Hence, we find that

(1) (1) −1/2 X 2πitng(x∗) In ∼ cn t e (H11) Appendix H: Asymptotic analysis for quantum x∗∈Σ oscillations (2) (2) −1 X 2πit(n g(x )+n g(x0 )) I ∼ c t e 1 ∗ 2 ∗ , n1,n2 n1,n2 0 x∗,x∗∈Σ In this appendix, we will derive the asymptotic form (H12) for quantum oscillations stated in Section VI C 3. Ab- I(3) ∼ · · · (H13) stracting out from the specific details, the situation is n1,n2,n3 33 where Σ is the set of solutions to g0(x) = 0. The os- with a quasiperiodic function of t with base frequencies cillatory part of each of these expressions is consistent ω(x∗) = 2πg(x∗), x∗ ∈ Σ.

[1] Elliott Lieb, Theodore Schultz, and Daniel Mattis, “Anomaly manifestation of Lieb-Schultz-Mattis theo- “Two soluble models of an antiferromagnetic chain,” rem and topological phases,” Phys. Rev. B 96, 195105 Ann. Phys. 16, 407 (1961). (2017), arXiv:1705.03892. [2] Masaki Oshikawa, Masanori Yamanaka, and Ian Af- [17] Chao-Ming Jian, Zhen Bi, and Cenke Xu, “Lieb- fleck, “Magnetization plateaus in spin chains: “Haldane Schultz-Mattis theorem and its generalizations from gap” for half-integer spins,” Phys. Rev. Lett. 78, 1984 the perspective of the symmetry-protected topo- (1997), arXiv:cond-mat/9610168. logical phase,” Phys. Rev. B 97, 054412 (2018), [3] Masaki Oshikawa, “Commensurability, excitation gap, arXiv:1705.00012. and topology in quantum many-particle systems on a [18] Max A. Metlitski and Ryan Thorngren, “Intrinsic and periodic lattice,” Phys. Rev. Lett. 84, 1535 (2000), emergent anomalies at deconfined critical points,” Phys. arXiv:cond-mat/9911137. Rev. B 98, 085140 (2018), arXiv:1707.07686. [4] M. B Hastings, “Sufficient conditions for topological [19] Yuan Yao and Masaki Oshikawa, “Generalized bound- order in insulators,” Europhys. Lett. 70, 824 (2005), ary condition applied to lieb-schultz-mattis-type ingap- arXiv:cond-mat/0411094. pabilities and many-body chern numbers,” Phys. Rev. [5] G. Misguich, C. Lhuillier, M. Mambrini, and P. Sindz- X 10, 031008 (2020), arXiv:1906.11662. ingre, “Degeneracy of the ground-state of antiferromag- [20] Xue-Yang Song, Yin-Chen He, Ashvin Vishwanath, and netic spin-1/2 Hamiltonians,” Eur. Phys. J. B 26, 167 Chong Wang, “Electric polarization as a nonquantized (2002), arXiv:cond-mat/0112360. topological response and boundary Luttinger theorem,” [6] J. M. Luttinger, “Fermi surface and some simple equi- arXiv:1909.08637. librium properties of a system of interacting fermions,” [21] P. Coleman, I. Paul, and J. Rech, “Sum rules and Phys. Rev. 119, 1153 (1960). Ward identities in the Kondo lattice,” Phys. Rev. B 72, [7] Masaki Oshikawa, “Topological approach to Luttinger’s 094430 (2005), arXiv:cond-mat/0503001. theorem and the Fermi surface of a kondo lattice,” Phys. [22] Stephen Powell, , and Hans Peter Rev. Lett. 84, 3370 (2000), arXiv:cond-mat/0002392. B¨uchler, “Depletion of the Bose-Einstein condensate in [8] T. Senthil, Subir Sachdev, and Matthias Vojta, “Frac- Bose-Fermi mixtures,” Phys. Rev. B 72, 024534 (2005), tionalized Fermi liquids,” Phys. Rev. Lett. 90, 216403 arXiv:cond-mat/0502299. (2003), arXiv:cond-mat/0209144. [23] Liza Huijse and Subir Sachdev, “Fermi surfaces and [9] T. Senthil, Matthias Vojta, and Subir Sachdev, “Weak gauge-gravity duality,” Phys. Rev. D 84, 026001 (2011), magnetism and non-Fermi liquids near heavy-fermion arXiv:1104.5022. critical points,” Phys. Rev. B 69, 035111 (2004), [24] M. Potthoff, “Non-perturbative construction of the arXiv:cond-mat/0305193. luttinger-ward functional,” Condens. Matt. Phys. 9, 557 [10] Arun Paramekanti and Ashvin Vishwanath, “Extend- (2006), cond-mat/0406671. ing Luttinger’s theorem to Z2 fractionalized phases of [25] Davide Gaiotto, Anton Kapustin, Nathan Seiberg, and matter,” Phys. Rev. B 70, 245118 (2004), arXiv:cond- Brian Willett, “Generalized global symmetries,” J. High mat/0406619. Energy Phys. 2015, 172 (2015), arXiv:1412.5148. [11] Parsa Bonderson, Meng Cheng, Kaushal Patel, and [26] Xie Chen, Zheng-Xin Liu, and Xiao-Gang Wen, “Two- Eugeniu Plamadeala, “Topological enrichment of Lut- dimensional symmetry-protected topological orders and tinger’s theorem,” arXiv:1601.07902. their protected gapless edge excitations,” Phys. Rev. B [12] Haruki Watanabe, Hoi Chun Po, Ashvin Vishwanath, 84, 235141 (2011), arXiv:1106.4752. and Michael Zaletel, “Filling constraints for spin-orbit [27] Dominic V. Else and Chetan Nayak, “Classifying coupled insulators in symmorphic and nonsymmorphic symmetry-protected topological phases through the crystals,” Proc. Natl. Acad. Sci. 112, 14551 (2015), anomalous action of the symmetry on the edge,” Phys. arXiv:1505.04193. Rev. B 90, 235137 (2014), arXiv:1409.5436. [13] Yuan-Ming Lu, Ying Ran, and Masaki Oshikawa, [28] G. ’t Hooft, “Naturalness, chiral symmetry, and spon- “Filling-enforced constraint on the quantized Hall con- taneous chiral symmetry breaking,” in NATO Advanced ductivity on a periodic lattice,” Ann. Phys. 413, 168060 Study Institute Seminar Series, Vol. 59 (1980) p. 135. (2020), arXiv:1705.09298. [29] Anton Kapustin and Ryan Thorngren, “Anomalous [14] Nick Bultinck and Meng Cheng, “Filling constraints discrete symmetries in three dimensions and group on fermionic topological order in zero magnetic field,” cohomology,” Phys. Rev. Lett. 112, 231602 (2014), Phys. Rev. B 98, 161119 (2018), arXiv:1808.00324. arXiv:1403.0617. [15] Meng Cheng, Michael Zaletel, Maissam Barkeshli, [30] Zheng-Cheng Gu and Xiao-Gang Wen, “Tensor- Ashvin Vishwanath, and Parsa Bonderson, “Trans- entanglement-filtering approach and lational symmetry and microscopic constraints on symmetry-protected topological order,” Phys. Rev. B symmetry-enriched topological phases: A view from 80, 155131 (2009), arXiv:0903.1069. the surface,” Phys. Rev. X 6, 041068 (2016), [31] T. Senthil, “Symmetry-protected topological phases of arXiv:1511.02263. quantum matter,” Annu. Rev. Condens. Matter Phys. [16] Gil Young Cho, Chang-Tse Hsieh, and Shinsei Ryu, 6, 299 (2015), arXiv:1405.4015. 34

[32] Xiao-Gang Wen, “Colloquium: Zoo of quantum- [48] C. Wang, A. C. Potter, and T. Senthil, “Classification topological phases of matter,” Rev. Mod. Phys. 89, of interacting electronic topological insulators in three 041004 (2017), arXiv:1610.03911. dimensions,” Science 343, 629 (2014), arXiv:1306.3238. [33] M. Z. Hasan and C. L. Kane, “Colloquium: Topo- [49] Chong Wang and T. Senthil, “Interacting fermionic logical insulators,” Rev. Mod. Phys. 82, 3045 (2010), topological insulators/superconductors in three arXiv:1002.3895. dimensions,” Phys. Rev. B 89, 195124 (2014), [34] F.D.M. Haldane, “Continuum dynamics of the 1-D arXiv:1401.1142. Heisenberg antiferromagnet: Identification with the [50] Anton Kapustin, “Symmetry protected topological O(3) nonlinear sigma model,” Phys. Lett. A 93, 464 phases, anomalies, and cobordisms: Beyond group co- (1983). homology,” arXiv:1403.1467. [35] F. D. M. Haldane, “Nonlinear field theory of large-spin [51] Max A. Metlitski, Lukasz Fidkowski, Xie Chen, and heisenberg antiferromagnets: Semiclassically quantized Ashvin Vishwanath, “Interaction effects on 3D topo- solitons of the one-dimensional easy-axis N´eelstate,” logical superconductors: surface topological order from Phys. Rev. Lett. 50, 1153 (1983). vortex condensation, the 16 fold way and fermionic [36] Frank Pollmann, Ari M. Turner, Erez Berg, and Masaki Kramers doublets,” arXiv:1406.3032. Oshikawa, “Entanglement spectrum of a topological [52] Anton Kapustin, Ryan Thorngren, Alex Turzillo, and phase in one dimension,” Phys. Rev. B 81, 064439 Zitao Wang, “Fermionic symmetry protected topologi- (2010), arXiv:0910.1811. cal phases and cobordisms,” J. High Energy Phys. 2015, [37] Xiao-Liang Qi, Taylor L. Hughes, and Shou-Cheng 1 (2015), arXiv:1406.7329. Zhang, “Topological field theory of time-reversal in- [53] Meng Cheng, Zhen Bi, Yi-Zhuang You, and Zheng- variant insulators,” Phys. Rev. B 78, 195424 (2008), Cheng Gu, “Classification of symmetry-protected arXiv:0802.3537. phases for interacting fermions in two dimensions,” [38] Frank Pollmann, Erez Berg, Ari M. Turner, and Masaki Phys. Rev. B 97, 205109 (2018), arXiv:1501.01313. Oshikawa, “Symmetry protection of topological phases [54] Alexei Kitaev, (2015), Institute for Pure Ap- in one-dimensional quantum spin systems,” Phys. Rev. plied Mathematics, UCLA. http://www.ipam.ucla. B 85, 075125 (2012), arXiv:0909.4059. edu/abstract/?tid=12389&pcode=STQ2015. [39] Xie Chen, Zheng-Cheng Gu, and Xiao-Gang Wen, [55] Edward Witten, “Fermion path integrals and topo- “Classification of gapped symmetric phases in one- logical phases,” Rev. Mod. Phys. 88, 035001 (2016), dimensional spin systems,” Phys. Rev. B 83, 035107 arXiv:1508.04715. (2011), arXiv:1008.3745. [56] Daniel S. Freed and Michael J. Hopkins, “Reflec- [40] Norbert Schuch, David P´erez-Garc´ıa, and Ignacio tion positivity and invertible topological phases,” Cirac, “Classifying quantum phases using matrix prod- arXiv:1604.06527. uct states and projected entangled pair states,” Phys. [57] Charles Zhaoxi Xiong, “Minimalist approach to the clas- Rev. B 84, 165139 (2011), arXiv:1010.3732. sification of symmetry protected topological phases,” [41] Xie Chen, Zheng-Cheng Gu, and Xiao-Gang Wen, J. Phys. A: Math. Theor. 51, 445001 (2018), “Complete classification of one-dimensional gapped arXiv:1701.00004. quantum phases in interacting spin systems,” Phys. [58] Anton Kapustin and Ryan Thorngren, “Fermionic SPT Rev. B 84, 235128 (2011), arXiv:1103.3323. phases in higher dimensions and bosonization,” J. High [42] Xie Chen, Zheng-Cheng Gu, Zheng-Xin Liu, and Xiao- Energy Phys. 2017, 80 (2017), arXiv:1701.08264. Gang Wen, “Symmetry protected topological orders and [59] Qing-Rui Wang and Zheng-Cheng Gu, “Towards a com- the group cohomology of their symmetry group,” Phys. plete classification of symmetry-protected topological Rev. B 87, 155114 (2013), arXiv:1106.4772. phases for interacting fermions in three dimensions and [43] Zheng-Cheng Gu and Xiao-Gang Wen, “Symmetry- a general group supercohomology theory,” Phys. Rev. protected topological orders for interacting fermions: X 8, 011055 (2018), arXiv:1703.10937. Fermionic topological nonlinear σ models and a spe- [60] Davide Gaiotto and Theo Johnson-Freyd, “Symme- cial group supercohomology theory,” Phys. Rev. B 90, try protected topological phases and generalized co- 115141 (2014), arXiv:1201.2648. homology,” J. High Energy Phys. 2019, 7 (2019), [44] Michael Levin and Zheng-Cheng Gu, “Braiding arXiv:1712.07950. statistics approach to symmetry-protected topolog- [61] Qing-Rui Wang and Zheng-Cheng Gu, “Construction ical phases,” Phys. Rev. B 86, 115109 (2012), and classification of symmetry-protected topological arXiv:1202.3120. phases in interacting fermion systems,” Phys. Rev. X [45] Yuan-Ming Lu and Ashvin Vishwanath, “Theory and 10, 031055 (2020), arXiv:1811.00536. classification of interacting integer topological phases [62] R. B. Laughlin, “Quantized Hall conductivity in two in two dimensions: A Chern-Simons approach,” Phys. dimensions,” Phys. Rev. B 23, 5632 (1981). Rev. B 86, 125119 (2012), arXiv:1205.3156. [63] F. D. M. Haldane, “Luttinger’s theorem and bosoniza- [46] Ashvin Vishwanath and T. Senthil, “Physics of three- tion of the Fermi surface,” in Proceedings of the Inter- dimensional bosonic topological insulators: Surface- national School of Physics “Enrico Fermi” (1992) p. 5, deconfined criticality and quantized magnetoelectric ef- arXiv:cond-mat/0505529. fect,” Phys. Rev. X 3, 011016 (2013), arXiv:1209.3058. [64] Andrew Pressley and Graeme Segal, Loop groups (Ox- [47] Lukasz Fidkowski, Xie Chen, and Ashvin Vish- ford University Press, Oxford). wanath, “Non-abelian topological order on the sur- [65] Ming-Che Chang and Qian Niu, “Berry phase, hyper- face of a 3d topological superconductor from an ex- orbits, and the Hofstadter spectrum: Semiclassical dy- actly solved model,” Phys. Rev. X 3, 041016 (2013), namics in magnetic Bloch bands,” Phys. Rev. B 53, arXiv:1305.5851. 7010 (1996), arXiv:cond-mat/9511014. 35

[66] Daniel G. Barci, Eduardo Fradkin, and Leonardo [85] Dam Thanh Son, “Is the a Ribeiro, “Bosonization of Fermi liquids in a weak Dirac particle?” Phys. Rev. X 5, 031027 (2015), magnetic field,” Phys. Rev. B 98, 155146 (2018), arXiv:1502.03446. arXiv:1805.05337. [86] Nathan Seiberg, T. Senthil, Chong Wang, and Edward [67] Siavash Golkar, Dung Xuan Nguyen, Matthew M. Witten, “A duality web in 2+1 dimensions and con- Roberts, and Dam Thanh Son, “Higher-spin theory densed matter physics,” Ann. Phys. 374, 395 (2016), of the magnetorotons,” Phys. Rev. Lett. 117, 216403 arXiv:1606.01989. (2016), arXiv:1602.08499. [87] Ajit C. Balram, Csaba T˝oke, and J. K. Jain, “Luttinger [68] T. Holstein, R. E. Norton, and P. Pincus, “de Haas-van theorem for the strongly correlated Fermi liquid of com- Alphen effect and the specific heat of an electron gas,” posite fermions,” Phys. Rev. Lett. 115, 186805 (2015), Phys. Rev. B 8, 2649 (1973). arXiv:1506.02747. [69] Patrick A. Lee, Naoto Nagaosa, and Xiao-Gang Wen, [88] Ajit C. Balram and J. K. Jain, “Fermi wave vector “Doping a Mott insulator: Physics of high-temperature for the partially spin-polarized composite-fermion Fermi ,” Rev. Mod. Phys. 78, 17 (2006), sea,” Phys. Rev. B 96, 235102 (2017), arXiv:1707.08623. arXiv:cond-mat/0410445. [89] B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida, [70] B. I. Halperin, Patrick A. Lee, and Nicholas Read, and J. Zaanen, “From quantum matter to high- “Theory of the half-filled Landau level,” Phys. Rev. B temperature superconductivity in copper oxides,” Na- 47, 7312 (1993). ture 518, 179 (2015), arXiv:1409.4673. [71] Joseph Polchinski, “Low-energy dynamics of the spinon- [90] Xiangang Wan, Ari M. Turner, Ashvin Vishwanath, gauge system,” Nucl. Phys. B 422, 617 (1994), and Sergey Y. Savrasov, “Topological semimetal and arXiv:cond-mat/9303037. Fermi-arc surface states in the electronic structure of [72] Chetan Nayak and Frank Wilczek, “Non-Fermi liquid pyrochlore iridates,” Phys. Rev. B 83, 205101 (2011). fixed point in 2 + 1 dimensions,” Nucl. Phys. B 417, [91] N. Read and Subir Sachdev, “Large-N expansion for 359 (1994), arXiv:cond-mat/9312086. frustrated quantum antiferromagnets,” Phys. Rev. Lett. [73] B. L. Altshuler, L. B. Ioffe, and A. J. Millis, “Low- 66, 1773 (1991). energy properties of fermions with singular interac- [92] X. G. Wen, “Mean-field theory of spin-liquid states with tions,” Phys. Rev. B 50, 14048 (1994), arXiv:cond- finite energy gap and topological orders,” Phys. Rev. B mat/9406024. 44, 2664 (1991). [74] Sung-Sik Lee, “Low-energy effective theory of Fermi sur- [93] T. Senthil and Matthew P. A. Fisher, “Z2 gauge the- face coupled with U(1) gauge field in 2 + 1 dimensions,” ory of electron fractionalization in strongly correlated Phys. Rev. B 80, 165102 (2009), arXiv:0905.4532. systems,” Phys. Rev. B 62, 7850 (2000), arXiv:cond- [75] Max A. Metlitski and Subir Sachdev, “Quantum phase mat/9910224. transitions of metals in two spatial dimensions. I. [94] Rodolfo A. Jalabert and Subir Sachdev, “Spontaneous Ising-nematic order,” Phys. Rev. B 82, 075127 (2010), alignment of frustrated bonds in an anisotropic, three- arXiv:1001.1153. dimensional Ising model,” Phys. Rev. B 44, 686 (1991). [76] David F. Mross, John McGreevy, Hong Liu, and [95] Subir Sachdev and Matthias Vojta, “Translational sym- T. Senthil, “Controlled expansion for certain non- metry breaking in two-dimensional antiferromagnets Fermi-liquid metals,” Phys. Rev. B 82, 045121 (2010), and superconductors,” in Proceedings of the Interna- arXiv:1003.0894. tional Workshop on Magnetic Excitations in Strongly [77] Yoni Schattner, Samuel Lederer, Steven A. Kivelson, Correlated Electrons, Hammatsu, Japan: August 19-22, and Erez Berg, “Ising nematic quantum critical point 1999, arXiv:cond-mat/9910231. in a metal: A Monte Carlo study,” Phys. Rev. X 6, [96] Pavel Etingof, Dmitri Nikshych, Victor Ostrik, and 031028 (2016), arXiv:1511.03282. with an appendix by Ehud Meir, “Fusion categories and [78] Denis Dalidovich and Sung-Sik Lee, “Perturbative non- homotopy theory,” arXiv:0909.3140. Fermi liquids from dimensional regularization,” Phys. [97] Maissam Barkeshli, Parsa Bonderson, Meng Cheng, Rev. B 88, 245106 (2013), arXiv:1307.3170. and Zhenghan Wang, “Symmetry fractionalization, de- [79] Sung-Sik Lee, “Recent developments in non-Fermi liq- fects, and gauging of topological phases,” Phys. Rev. B uid theory,” Annu. Rev. Condens. Matter Phys. 9, 227 100, 115147 (2019), arXiv:1410.4540. (2018), arXiv:1703.08172. [98] Francesco Benini, Clay C´ordova, and Po-Shen Hsin, [80] Max A. Metlitski, David F. Mross, Subir Sachdev, and “On 2-group global symmetries and their anomalies,” J. T. Senthil, “Cooper pairing in non-Fermi liquids,” Phys. High Energy Phys. 2019, 118 (2019), arXiv:1803.09336. Rev. B 91, 115111 (2015), arXiv:1403.3694. [99] Po-Shen Hsin and Alex Turzillo, “Symmetry-enriched [81] T. Senthil, “Critical Fermi surfaces and non-Fermi quantum spin liquids in (3 + 1)d,” J. High Energy Phys. liquid metals,” Phys. Rev. B 78, 035103 (2008), 2020, 22 (2020), arXiv:1904.11550. arXiv:0803.4009. [100] John C. Baez and Aaron D. Lauda, “Higher-dimensional [82] D. Shoenberg, Magnetic Oscillations in Metals (Cam- algebra V: 2-groups,” arXiv:math.QA/0307200. bridge University Press, Cambridge, 1984). [101] Sergei Gukov and Anton Kapustin, “Topological quan- [83] Jing-Yuan Chen and Dam Thanh Son, “Berry tum field theory, nonlocal operators, and gapped phases Fermi liquid theory,” Ann. Phys. 377, 345 (2017), of gauge theories,” arXiv:1307.4793. arXiv:1604.07857. [102] Anton Kapustin and Ryan Thorngren, “Higher [84] Dam Thanh Son and Naoki Yamamoto, “Berry curva- symmetry and gapped phases of gauge theories,” ture, triangle anomalies, and the chiral magnetic effect arXiv:1309.4721. in Fermi liquids,” Phys. Rev. Lett. 109, 181602 (2012), [103] C. Lannert, Matthew P. A. Fisher, and T. Senthil, arXiv:1203.2697. “Quantum confinement transition in a d-wave supercon- 36

ductor,” Phys. Rev. B 63, 134510 (2001), arXiv:cond- 2012, 64 (2012), arXiv:1203.5388. mat/0007002. [114] W. O. Putikka, M. U. Luchini, and R. R. P. [104] Leon Balents, Lorenz Bartosch, Anton Burkov, Subir Singh, “Violation of Luttinger’s Theorem in the Two- Sachdev, and Krishnendu Sengupta, “Putting compet- Dimensional t- J Model,” Phys. Rev. Lett. 81, 2966 ing orders in their place near the Mott transition,” Phys. (1998), arXiv:cond-mat/9803140. Rev. B 71, 144508 (2005), arXiv:cond-mat/0408329. [115] Emmanuele Cappelluti and Roland Zeyher, “Violation [105] Igor Dzyaloshinskii, “Some consequences of the Lut- of luttinger’s theorem in strongly correlated electronic tinger theorem: The Luttinger surfaces in non-Fermi systems within a 1/N expansion,” Int. J. Mod Phys B liquids and Mott insulators,” Phys. Rev. B 68, 085113 13, 2607 (1999), arXiv:cond-mat/9906349. (2003). [116] Dominic V. Else and T. Senthil, “Strange metals as er- [106] Tudor D. Stanescu and Gabriel Kotliar, “Fermi arcs and satz Fermi liquids,” arXiv:2010.10523. hidden zeros of the Green function in the pseudogap [117] Alexei Kitaev, “Periodic table for topological insula- state,” Phys. Rev. B 74, 125110 (2006), arXiv:cond- tors,” AIP Conference Proceedings 1134, 22 (2009), mat/0508302. arXiv:0901.2686. [107] Kai-Yu Yang, T. M. Rice, and Fu-Chun Zhang, “Phe- [118] Ryan Thorngren and Dominic V. Else, “Gauging spa- nomenological theory of the pseudogap state,” Phys. tial symmetries and the classification of topological Rev. B 73, 174501 (2006), arXiv:cond-mat/0602164. crystalline phases,” Phys. Rev. X 8, 011040 (2018), [108] A. Rosch, “Breakdown of Luttinger’s theorem in two- arXiv:1612.00846. orbital Mott insulators,” Eur. Phys. J. B 59, 495 (2007), [119] Dominic V. Else, Wen Wei Ho, and Philipp T. arXiv:cond-mat/0602656. Dumitrescu, “Long-lived interacting phases of matter [109] Kiaran B. Dave, Philip W. Phillips, and Charles L. protected by multiple time-translation symmetries in Kane, “Absence of Luttinger’s theorem due to zeros quasiperiodically driven systems,” Phys. Rev. X 10, in the single-particle green function,” Phys. Rev. Lett. 021032 (2020), arXiv:1910.03584. 110, 090403 (2013), arXiv:1207.4201. [120] C. Duval, Z. Horv´ath,P. A. Horv´athy, L. Martini, and [110] S. A. Hartnoll, D. M. Hofman, and A. Tavanfar, “Holo- P. C. Stichel, “Berry phase correction to electron density graphically smeared Fermi surface: Quantum oscilla- and “exotic” dynamics,” Mod. Phys. Lett. B 20, 373 tions and Luttinger count in electron stars,” Europhys. (2006), arXiv:cond-mat/0506051. Lett. 95, 31002 (2011), arXiv:1011.2502. [121] John M. Lee, Introduction to smooth manifolds [111] Sean A. Hartnoll, “Horizons, holography and condensed (Springer, New York, 2013). matter,” in Black Holes in Higher Dimensions, edited [122] https://mathoverflow.net/questions/361149/does- by Gary T. Horowitz (Cambridge University Press, New a-compact-lie-group-have-finitely-many- York, 2012) pp. 387–419, arXiv:1106.4324. conjugacy-classes-of-maximal-abelian. [112] Nabil Iqbal and Hong Liu, “Luttinger’s theorem, su- [123] Deane Montgomery and Leo Zippin, “A theorem of Lie perfluid vortices and holography,” Classical Quantum groups,” Bulletin of the American Mathematical Society Gravity 29, 194004 (2012), arXiv:1112.3671. 48, 448 (1942). [113] Koji Hashimoto and Norihiro Iizuka, “A comment on [124] Ryan Thorngren, “Anomalies and bosonization,” holographic Luttinger theorem,” J. High Energy Phys. arXiv:1810.04414.