diff --git "a/DtAzT4oBgHgl3EQfwv4v/content/tmp_files/2301.01726v1.pdf.txt" "b/DtAzT4oBgHgl3EQfwv4v/content/tmp_files/2301.01726v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/DtAzT4oBgHgl3EQfwv4v/content/tmp_files/2301.01726v1.pdf.txt" @@ -0,0 +1,3351 @@ +Anisotropic Quantum Hall Droplets +Blagoje Oblak,1 Bastien Lapierre,2 Per Moosavi,3 Jean-Marie Stéphan,4 and Benoit Estienne5 +1CPHT, CNRS, École Polytechnique, IP Paris, F-91128 Palaiseau, France +2Department of Physics, University of Zürich, Winterthurerstrasse 190, 8057 Zürich, Switzerland +3Institute for Theoretical Physics, ETH Zurich, Wolfgang-Pauli-Strasse 27, 8093 Zürich, Switzerland +4Univ Lyon, CNRS, Université Claude Bernard Lyon 1, +Institut Camille Jordan, UMR5208, F-69622 Villeurbanne, France +5Sorbonne Université, CNRS, Laboratoire de Physique Théorique et Hautes Energies, LPTHE, F-75005 Paris, France +(Dated: January 4, 2023) +We study two-dimensional (2D) anisotropic droplets of non-interacting electrons in the lowest +Landau level, confined by trapping potentials whose level curves have an arbitrary shape at large +distances. Using semiclassical methods, we show that energy eigenstates are localized on equipoten- +tials of the trap, with angle-dependent local widths and heights. We exploit this one-particle insight +to deduce explicit formulas for many-body observables in the thermodynamic limit. For instance, +the droplet’s density falls off at the boundary with an angle-dependent width inherited from that of +the underlying wave functions, while the many-body current is localized on the edge, to which it is +tangent. Correlations along the edge are long-ranged, in accordance with the system’s low-energy +edge modes which are described by a free chiral conformal field theory in terms of the angle variable +of the trapping potential. These results are likely to be observable in solid-state systems or quantum +simulators of 2D electron gases with a high degree of control on the confining potential. +CONTENTS +I. Introduction +1 +II. Setup and main results +2 +III. Anisotropic states from area-preserving maps +4 +IV. Edge-deformed anisotropic traps +6 +V. Many-body observables +10 +VI. Conclusion and outlook +14 +Acknowledgments +14 +A. Isotropic droplets +15 +B. Semiclassical expansion of P V P +15 +C. The transport equation +16 +References +19 +I. +INTRODUCTION +Quantum Hall (QH) droplets are mesoscopic two- +dimensional (2D) electron gases placed in a strong per- +pendicular magnetic field and confined by some electro- +static potential: see Fig. 1. They lie at the heart of the +QH effect [1–3] and provide a key benchmark for topolog- +ical phases of matter as a whole. In practice, however, +the vast majority of detailed analytical studies of QH +droplets are limited to highly symmetric cases, typically +involving isotropic traps or harmonic potentials that are +translation-invariant in one direction [4, 5]. This is espe- +cially troubling as far as edge properties are concerned, +since these are sensitive to the shape of the trap and +determine the system’s low-energy excitations [6–10]. +The goal of this paper is to address this lack of analyt- +ical results by predicting universal aspects of many-body +observables near the edge of essentially any anisotropic +droplet. We achieve this by providing general, explicit, +one-line formulas for the density, current, and correla- +tions in the regime of strong magnetic fields. We also +study the corresponding low-energy edge modes, which +are described by a free-fermion chiral conformal field the- +ory (CFT) whose Fermi velocity is constant provided dis- +tances along the boundary are measured by the canonical +angle variable determined by the potential. These pre- +dictions are likely to be observable thanks to direct local +imaging techniques in condensed matter systems [11–16] +or quantum simulators [17–24]. +This is not the first time such questions appear in the +literature. Indeed, random potentials with no symme- +tries are essential to model disorder, whose importance +for the robustness of QH physics is hard to overstate [25– +27]. An especially relevant series of works in that con- +text is [28, 29], which study the density and current of +QH droplets with arbitrary potentials, at finite temper- +ature, generally including Landau level mixing, in the +semiclassical limit of strong magnetic fields and weak +traps [26, 27]. However, the coherent states used in these +references do not allow for any resolution at the single- +particle level, precluding the computation of low-energy +dynamics and long-range correlations along the bound- +ary. Our objective here is instead to find explicit wave +functions that depend on the shape of the edge, and use +that as a starting point for many-body objects. +Regarding electronic edge correlations, similar ques- +tions have been addressed in the context of classical 2D +arXiv:2301.01726v1 [cond-mat.mes-hall] 4 Jan 2023 + +2 +Energy +Fermi +energy +Figure 1. +2D electron droplet (shaded area) placed in a +strong perpendicular magnetic field and confined by a typical +anisotropic edge-deformed potential well (10). At leading or- +der in the thermodynamic limit, the droplet’s boundary (thick +black curve) coincides with the equipotential of the trap at the +Fermi energy. +Coulomb gases, where holomorphic methods play a key +role [30–32]. More broadly, the results put forward here +may be seen as microscopic, first-principles derivations +of quantities that are normally studied within less con- +trolled approximation schemes in the geometry of the QH +effect [33–39]. Our hope is thus to build a bridge between +these theoretical works and concrete observations that +may soon be accessible in tabletop experiments with a +high degree of control on the confining potential [23, 24]. +Here is the plan of the paper. To begin, Sec. II sum- +marizes our methods and results, avoiding all techni- +cal details. The next two sections are devoted to one- +body physics in the lowest Landau level: Sec. III first +discusses generalities on semiclassical holomorphic wave +functions, while Sec. IV presents a detailed computation +of the semiclassical energy spectrum in a class of ‘edge- +deformed’ potentials of particular interest. This finally +leads to Sec. V, where we investigate many-body densi- +ties, currents, correlations, and low-energy edge modes in +anisotropic traps. We conclude in Sec. VI by discussing +several follow-ups and open questions. To streamline the +text, non-essential details are deferred to Apps. A–C. +II. +SETUP AND MAIN RESULTS +This section summarizes our methods and key results. +To start, we describe the general setup: a semiclassi- +cal limit (strong magnetic field, small magnetic length) +in the lowest Landau level (LLL) [26, 27, 40]. We then +introduce edge-deformed potentials and present their ap- +proximate energy eigenstates, before finally giving simple +formulas for the corresponding local many-body observ- +ables and dynamics in the vicinity of the edge. Some of +these findings are depicted in Fig. 2. +A. +Semi-classical limit in the LLL +This work concerns spin-polarized non-interacting elec- +trons of mass M and charge q in a 2D plane. +Each +electron is governed by a Landau Hamiltonian with an +anisotropic potential V (x), +H1-body = +1 +2M (p − qA)2 + V (x), +(1) +where A is the vector potential of the magnetic field +B = dA. (We view A as a one-form, which simplifies +some notation but is otherwise inconsequential.) We shall +assume that V (x) is ‘monotonous’, by which we mean +that it has a unique global minimum away from which it +grows monotonously, but it is otherwise general. Conse- +quently, the level curves or ‘equipotentials’ of V (x) are +nested and take the form shown in Fig. 2(a). We assume +throughout that the potential is weak relative to the mag- +netic field [26–29, 41, 42], and that it is nearly constant +on length scales comparable to the magnetic length +ℓ2 ≡ ℏ +qB > 0. +(2) +In that regime, the potential is a small perturbation of +the pure Landau Hamiltonian ∝ (p−qA)2 and the eigen- +states of (1) are expected to be well-approximated by +wave functions in the LLL. For instance, if the potential +V (x) = V0(r2/2) is isotropic, the eigenstates of (1) have +some definite angular momentum and reduce at strong +B to standard LLL wave functions in symmetric gauge: +φm(x) = +1 +√ +2πℓ2 +zm +√ +m! +e−|z|2/2, +(3) +where m ≥ 0 is an integer angular momentum and we +have introduced the dimensionless complex coordinate +z ≡ x + iy +√ +2ℓ . +(4) +Each wave function (3) reaches its maximum on the cir- +cle |z| = √m, away from which it decays in a Gaussian +manner within a magnetic length. Our goal will be to ob- +tain similar approximate eigenstates in anisotropic traps, +using the squared magnetic length (2) as a small expan- +sion parameter [43]. Equivalently, we shall carry out a +semiclassical (small ℏ), high field expansion (large B). +In practice, the projection to the LLL is implemented +thanks to the (one-body) operator P ≡ �∞ +m=0 |φm⟩⟨φm|, +whose kernel can be read off from the wave functions (3): +⟨z, ¯z|P|w, ¯w⟩ = +1 +2πℓ2 e−(|z|2+|w|2)/2 ez ¯ +w. +(5) +This is manifestly Gaussian and reduces to a delta func- +tion in the formal semiclassical limit ℓ → 0. At small but +finite ℓ, the projection (5) makes space non-commutative +in the sense that the LLL-projected position operators +(x, y) satisfy the Heisenberg algebra +[PxP, PyP] = iℓ2. +(6) + +3 +��� +��� +��� +x +x +x +y +y +y +√ +N +− +√ +N +− +√ +N +√ +N +(a) +(b) +(c) +x +x +x +y +y +y +ℓ +√ +2N +−ℓ +√ +2N +−ℓ +√ +2N +ℓ +√ +2N +Figure 2. (a) A planar plot of the many-body density (14) along with several equipotentials (dashed lines), for a droplet of +N = 100 electrons confined by the edge-deformed trap of Fig. 1. The constancy of the bulk density and its fall at the boundary +are manifest. (b) The current’s norm (15) for the same droplet, together with the edge (black line) on which it is localized. +(c) The correlation function (16) for the same droplet, seen as a function of x2 = (x, y) with x1 = (ℓ +� +Nλ(0), 0) denoted by a +cross and fixed at the edge. Long-range correlations along the boundary are clearly visible. +One can thus think of the plane R2 as a ‘phase space’ +whose canonical variables are x, y. This interpretation +pervades much of the QH literature [40, 44–53] and will +similarly affect our discussion. +Indeed, projecting the +Hamiltonian (1) to the LLL and looking for its spectrum +leads to the eigenvalue equation +P V P|ψ⟩ = E|ψ⟩, +(7) +where the unknowns are the energy E and the quantum +state |ψ⟩ ∈ LLL. Note that the kinetic term of (1) has +disappeared in (7): the potential itself plays the role of +an effective Hamiltonian in the non-commutative phase +space (x, y). +Exact solutions of Eq. (7) are generally out of reach, +so one has to resort to approximations. The semiclassical +one that we shall use is well known in the QH context +[26–29, 54, 55]. Accordingly, we will look for solutions of +Eq. (7) labeled by a large quantum number m ∈ N, seen +as a generalization of angular momentum. +This large +m limit is accompanied by a small ℓ limit, such that +the area 2πℓ2m remains finite. In that regime, the mth +eigenstate is approximately Gaussian and localized on an +equipotential γm of V (x), enclosing a quantized area such +that the Bohr-Sommerfeld condition holds: +� +γm +x dy = 2πℓ2 m. +(8) +Equivalently, the flux of the magnetic field through the +area enclosed by γm is m times the flux quantum. The +energy of the mth state is then +Em = E0 +m + ℓ2E1 +m + O(ℓ4), +(9) +where E0 +m = V (γm) is the leading classical approxima- +tion and the quantum correction E1 +m involves the Lapla- +cian of the potential and the curvature of the equipo- +tential γm [54, 55]. Note that the more familiar Wentzel- +Kramers-Brillouin (WKB) approximation of 1D quantum +mechanics [56] includes (topological) Maslov corrections +on the right-hand side of (8); we will find similar correc- +tions below, although their interpretation as topological +invariants is prevented by a subtle distinction between +real and Kählerian polarizations in geometric quantiza- +tion [54, 55]. +B. +One-body results +The semiclassical limit just outlined applies to any +(monotonous) weak potential. In practice, our main con- +cern is the physics of QH droplets near the edge, where +the details of the bulk potential are irrelevant. +Most +of our explicit results will therefore be given for ‘edge- +deformed’ potentials, obtained as follows. Consider any +monotonously increasing function V0(t) for t ≥ 0, and let +λ(ϕ) be any strictly positive 2π-periodic function of the +angle ϕ ∈ [0, 2π); normalize λ so that +� +dϕ λ(ϕ) = 4π, +where we write +� +dϕ as a shorthand for +� 2π +0 +dϕ. Then +adopt polar coordinates in the plane such that x + iy = +r eiϕ and define the potential +V (r, ϕ) = V0 +� r2 +λ(ϕ) +� +. +(10) +We refer to this as an edge-deformed trap because it re- +sults from a deformation r2 �→ r2/λ(ϕ) that changes the +shape of the boundary of isotropic droplets in a finite +and smooth way, even in the thermodynamic limit where +the droplet’s area goes to infinity [57]. In fact, the cor- +responding infinitesimal transformations are expected to + +1 +1 +- +- +1 +1 +- +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +-4 +be conformal transformations of the edge CFT [58–63]. +The class of potentials (10) is thus exhaustive, at least +as far as edge effects are concerned. +The traps (10) turn out to allow for explicit calcula- +tions of the semiclassical energy spectrum, generalizing +the known isotropic results reviewed in App. A. Indeed, +we show in Sec. IV that the corresponding eigenfunc- +tions, solving (7) in the LLL, are Gaussians localized on +equipotentials r = ℓ +� +mλ(ϕ) at large quantum numbers +m. They can be written in polar coordinates as +ψm(x) ∼ +eiΘm(x) +� +2πℓ2σ(ϕ) +e−a2/σ(ϕ)2 +(2πm)1/4 , +(11) +where Θm(x) is a position-dependent phase, a ≡ +� +r − +ℓ +� +mλ(ϕ) +�� +ℓ +� +λ(ϕ) is a dimensionless radial coordinate +that measures the distance from the equipotential, and +σ(ϕ) ≡ +� +2 +λ(ϕ) +� +1 + +� λ′(ϕ) +2λ(ϕ) +�2 +(12) +is an angle-dependent width. We stress that this exhibits +the expected ‘quantum smearing’ of wave functions in a +strong but finite magnetic field [28, 29], which would be +missed by the leading classical approximations ℓ2 = 0. +As for the energy of the state (11), its expansion (9) +up to neglected O(ℓ4) contributions is +Em ∼ V (γm) + ℓ2 +2 Ωm +� +1 + +� +1 + Γm +Ωm +� � dϕ +4π λ(ϕ)σ(ϕ)2 +� +, +(13) +where V (γm) = V0 +� +ℓ2m +� +is the leading term, while +the first quantum correction involves derivatives Ωm ≡ +V ′ +0(ℓ2m) > 0 and Γm ≡ ℓ2m V ′′ +0 (ℓ2m). Each mth energy +is thus determined by the potential and its derivatives at +an equipotential that satisfies the quantization condition +(8), in accordance with general theorems for holomorphic +WKB theory [54, 55]. +C. +Many-body results +Now consider the ground state of a large number +N ≫ 1 of free spin-polarized electrons, each governed by +the single-particle Hamiltonian (1). This ground state +is a Slater determinant of wave functions whose large m +behavior is the Gaussian (11). As we show in Sec. V, +the corresponding many-body density, current, correla- +tions, and low-energy effective action can all be written +in closed form in terms of λ(ϕ) and the number N of +fermions. The density ρ(x) = �N−1 +m=0 |ψm(x)|2 thus sat- +isfies the bulk behavior ρ ∼ +1 +2πℓ2 , while its form near the +edge is given by a complementary error function: +ρ(r, ϕ) ∼ +1 +4πℓ2 erfc +�√ +2 a +σ(ϕ) +� +, +(14) +where a ≡ +� +r − ℓ +� +Nλ(ϕ) +�� +ℓ +� +λ(ϕ) is again a dimen- +sionless radial coordinate, now measuring the distance +from the edge at redge = ℓ +� +Nλ(ϕ). +As a result, the +ground state forms a star-shaped droplet whose bound- +ary has an angle-dependent width (12) inherited from +that of one-body wave functions. Turning to the current +J = �N−1 +m=0 +1 +2i(ψ∗ +mdψm − ψmdψ∗ +m − 2iqA|ψm|2), we write +it as a one-form in polar coordinates to find +J(r, ϕ) ∼ − +exp +� +− 2a2 +σ(ϕ)2 +� +(2πℓ2)3/2σ(ϕ) +� +ℓ +√ +N dϕ + +λ′(ϕ) +2λ(ϕ)3/2 dr +� +. +(15) +This is localized on the edge and tangent to it, miss- +ing the bulk behavior Ji ∝ εij∂jV as expected in the +LLL [41, 42]. Finally, the two-point correlation function +C(x1, x2) = �N−1 +m=0 ψ∗ +m(x1)ψm(x2) behaves near the edge +as +C(x1, x2) ∼ +eiΘN(x1,x2) +4πℓ2� +σ(ϕ1)σ(ϕ2) +i exp +� +− +a2 +σ(ϕ1)2 − +b2 +σ(ϕ2)2 +� +√ +2πN sin +�� ϕ1 +ϕ2 +dθ +4 λ(θ) +� +(16) +with a ≡ +� +|x1| − ℓ +� +Nλ(ϕ1) +�� +ℓ +� +λ(ϕ1) and similarly for +b in polar coordinates (|x1|, ϕ1) and (|x2|, ϕ2), respec- +tively, while ΘN(x1, x2) is a complicated overall phase. +Note again the Gaussian localization at the edge, as well +as the long-range correlator ∝ sin(...)−1 typical of gapless +fermions. Indeed, we eventually confirm that the under- +lying low-energy edge modes are described by a chiral +CFT of free fermions: see the action functional (68) be- +low. The corresponding Fermi velocity is constant along +the boundary when measured in terms of the angle vari- +able of the potential (10), namely θ(ϕ) ≡ +� ϕ +0 dϕ′ λ(ϕ′)/2. +III. +ANISOTROPIC STATES FROM +AREA-PRESERVING MAPS +This section presents the WKB ansatz [see (20)] that +forms the basis of all our later considerations. The struc- +ture is ultimately quite simple: given a monotonous po- +tential V (x), we pick one of its equipotentials, γm, with +quantized area (8). We then build a wave function with +winding m, perfectly localized on γm, and finally project +it to the LLL using the operator (5). General theorems +[54, 55] ensure that LLL-projected eigenstates satisfying +(7) can indeed be built in this way. The detailed applica- +tion of this method to edge-deformed traps (10) is given +in Sec. IV. +A note: what follows relies on the mathematics of area- +preserving diffeomorphisms, which is not reviewed in de- +tail. We refer instead to [57] for an introduction whose +language is similar to that adopted here. For more gen- +eral discussions in the symplectic context, see [64, 65]. +A. +Potentials in action-angle variables +Let us be more precise about the geometry of the +setup, remaining at the classical level for now. We pick + +5 +a smooth potential V (x) and assume as in Sec. II that it +is monotonous. Its unique global minimum is thus sur- +rounded by nested level curves, and one can always find +an area-preserving deformation of the plane that sends +each equipotential on a circle [64]. In other words, one +can find an invertible smooth map F : R2 → R2 with +unit Jacobian such that +V +� +F(x) +� += V0(r2/2), +(17) +where the trap on the right-hand side is isotropic (it only +depends on r = |x|). If F is the identity (or a rotation +around the origin), then V was isotropic to begin with +and its eigenstates satisfying (7) are the standard wave +functions (3) with definite angular momentum. In the +more general case of arbitrary V , Eq. (17) suggests using +F to map the eigenstates (3) on those corresponding to +our general V (x). +The existence of F in (17) is guaranteed by the mono- +tonicity of V , and is equivalent to the existence of glob- +ally well-defined action-angle variables. In fact, we can +use this to write F in a more explicit form that will be +useful below. Let therefore (ℓ2K, θ) be action-angle co- +ordinates for the potential V (x) [66], where K ≥ 0 is +dimensionless and θ ∈ [0, 2π) is a genuine angle. They +are normalized so that ℓ2dK ∧ dθ = dx ∧ dy, which is +to say that their Poisson bracket reads {ℓ2K, θ} = ℓ2 in +terms of the phase space (x, y) with bracket (6). Then +the map (x, y) �→ (ℓ2K, θ) is an area-preserving diffeo- +morphism in terms of which V (x) = V0(ℓ2K(x)) is in- +variant under rotations of θ. To be specific, write these +coordinates as functions K(x, y) and θ(x, y) and let the +inverse be x = F(K, θ) and y = G(K, θ) for some func- +tions (F, G); this inverse is nothing but the deformation +F in (17). In other words, knowing the action-angle vari- +ables of a potential V allows us to map it on its (unique) +isotropic cousin V0, which in turn can be used to relate +the corresponding anisotropic eigenstates to those in (3). +It should be clear that these considerations apply to +any monotonous anisotropic trap, in which case one +generally encounters intricate area-preserving maps with +complicated action-angle variables. In Sec. IV, we will +argue that most of these difficulties wash away when +focusing on edge physics, whereupon the only relevant +maps are the ‘edge deformations’ mentioned below Eq. +(10). For now, we remain general and turn to quantum +considerations. +B. +Anisotropic eigenstates +Using the action-angle variables (ℓ2K, θ) for V (x), +one can concretize the statements around Eqs. (8)–(9) +into formulas and eventually obtain anisotropic eigen- +functions that satisfy (7). Indeed, the Bohr-Sommerfeld +quantization condition (8) implies that the equipotential +γm is the set of points in R2 where K = m. Now consider +the following quantum state, perfectly localized on γm: +|Ψm⟩ ≡ 2πℓ2 +� +dθ n(θ) eimθ��F(m, θ), G(m, θ) +� +, +(18) +where the normalization 2πℓ2 is included for later conve- +nience, the ‘wave function’ ⟨x|F(m, θ), G(m, θ)⟩ = δ2� +x− +F(m, θ) +� +is a delta function, and n(θ) is some complex +periodic function. The latter does not wind upon com- +pleting one turn in the plane along the equipotential, +meaning that all the winding of (18) is encoded in the +phase eimθ. +We stress that (18) is analogous to the standard WKB +ansatz ψ(x) ∼ eiS0(x)/ℏeiS1(x) in 1D. Indeed, the phase +eimθ is the leading classical contribution eiS0/ℏ for m ≫ 1, +corresponding to the ‘geometrical optics’ approximation +of the wave function, while n(θ) is the ‘physical optics’ +quantum correction eiS1 that eventually needs to satisfy a +transport equation in order for the Schrödinger equation +to hold [56]. The only difference lies in the interpretation +of areas in the plane as values of an ‘action’, which ul- +timately stems from the non-commutative geometry (6) +of LLL physics. Note that n(θ) is the only unknown in +(18). Indeed, most of the WKB method below will con- +cern the derivation of a transport equation for n(θ) from +the requirement that (7) be satisfied. +Starting from Eq. (18), it is straightforward to build a +state in the LLL thanks to the projector (5): denoting +ψm(z, ¯z) ≡ ⟨z, ¯z|P|Ψm⟩, +(19) +one finds the wave function +ψm(z, ¯z) = e−|z|2/2 +� +dθ n(θ) eimθ +× e−[F (m,θ)2+G(m,θ)2]/4ℓ2 ez[F (m,θ)−iG(m,θ)]/ +√ +2ℓ. +(20) +This is manifestly of the form e−|z|2/2 times a holomor- +phic function that depends on the action variable ℓ2m +and the uniformizing map F of Eq. (17). It will be our +starting point for the semiclassical solution of the eigen- +value equation (7). +As a consistency check, note that (20) simplifies for +isotropic potentials. In that case, action-angle variables +are just polar coordinates ℓ2K = r2/2 and θ = ϕ, and +the map in (17) is F(x) = x, merely implementing a +change from polar to Cartesian coordinates: F(m, θ) = +ℓ +√ +2m cos(θ) and G(m, θ) = ℓ +√ +2m sin(θ). One can then +verify that (20) with n(θ) = const coincides (up to nor- +malization) with the standard LLL wave function (3). +Similarly to that case, any projected wave function (20) +reaches its maximum on the equipotential γm and is ap- +proximately Gaussian close to it, as ensured by the kernel +(5). This will be confirmed explicitly in Sec. IV for edge +deformations. + +6 +C. +Expanding the eigenvalue equation +None of what we wrote so far involves a manifest semi- +classical expansion: it is hidden in the eigenvalue equa- +tion (7) and the function n(θ) in (20), since n(θ) should +be expanded as a power series n(θ) = n0(θ) + ℓ2n1(θ) + +O(ℓ4) (as before, there are no odd powers of ℓ since ℓ2 ∝ ℏ +is really the semiclassical parameter). +It is therefore +worth anticipating the first few terms of the semiclas- +sical approximation of (7). We stress that the expansion +below will eventually be limited to the leading order of +the transport equation, so that only n0(θ) will be calcu- +lated in the end. In principle, one could of course push +the expansion to higher orders for more detailed results. +The semiclassical expansion of the right-hand side of +(7) is clear: it is given by the large m, small ℓ2 expan- +sion of the projected wave function (20), including an +expansion of n(θ). As for the energy, its expansion was +written in (9). The left-hand side of (7) is more subtle, as +its semiclassical expansion involves that of the operator +P V P. The latter is a ‘Berezin-Toeplitz operator’ [54, 55] +that will play an important role for edge-deformed po- +tentials, so we now explain its expansion in some detail. +First, given Cartesian coordinates (x, y), express the po- +tential in complex coordinates (4) as V (x, y) ≡ V(z, ¯z) +for some function V(z, ¯w) which is holomorphic in its first +argument and anti-holomorphic in the second. Then re- +calling that P is the LLL projector with kernel (5), one +finds +⟨z, ¯z|P V P|w, ¯w⟩ = +1 +2πℓ2 e− 1 +2 (|z|2+|w|2) +× +� +R2 du dv V (u, v) e−|X|2+z ¯ +X+ ¯ +wX +(21) +with X ≡ (u + iv)/ +√ +2ℓ similarly to (4). Our task is to +expand the integral on the right-hand side in the semi- +classical limit. The key is to assume that the potential +varies slowly on the scale of the magnetic length [26–29], +i.e. we choose once and for all a smooth potential V (x), +independent of ℓ, and let ℓ be small. In that regime, the +integrals in (21) are approximately Gaussian, which gives +(see App. A 2) +⟨z, ¯z|P V P|w, ¯w⟩ +ℓ≪1 +∼ +1 +2πℓ2 e− |z−w|2 +2 +e +z ¯ +w−¯zw +2 +× +� +V(z, ¯w) + ℓ2 +2 (∇2V )(z, ¯w) +� +(22) +where +(∇2V )(z, ¯w) +is +the +bicomplex +function +that +corresponds to the Laplacian of the potential, +i.e. +(∇2V )(z, ¯w) = +4 +2ℓ2 ∂z∂ ¯ +wV. This is the standard semiclas- +sical expansion of a Berezin-Toeplitz operator [54, 55]. +Note the general structure: the entire P V P operator +boils down to P itself, with kernel (5), multiplied by a +function that coincides with V at leading order but also +includes quantum corrections. In the ‘zoomed-out’ limit +where the kernel of P is a delta function, the first term of +(22) becomes V(z, ¯z)δ2(z − w, ¯z − ¯w) as expected. More- +over, for harmonic potentials, the truncated expression +(22) is actually exact since the next term ∇4V and all +subsequent ones vanish. This agrees with the common +lore that ‘WKB is exact for quadratic Hamiltonians’. +IV. +EDGE-DEFORMED ANISOTROPIC TRAPS +Here we apply the WKB ansatz of Sec. III to poten- +tials (10) with scale-invariant level curves, obtained by +acting with edge deformations [57] on an isotropic trap. +As we explain below, these are the most general traps +one expects to find close to the edge of star-shaped QH +droplets. +The plan is as follows. +First, we introduce +edge deformations and give a few examples for later ref- +erence. Second, we apply Eq. (7) to edge-deformed traps +and expand it in the classical limit (large m, small ℓ2 +with ℓ2m = O(1) kept fixed). We keep track of all terms +up to order O(ℓ2), so as to capture the leading part of +the transport equation for the function n(θ) in (18)–(20). +This eventually yields an explicit energy spectrum [see +(37)] along with approximately Gaussian eigenfunctions +[see (43)]. Lastly, we conclude with a consistency check +by showing that our wave functions reproduce the asymp- +totic (large m) form of the known LLL-projected spec- +trum for anisotropic harmonic traps [67–71]. +A. +Edge deformations +We have seen in Sec. III that area-preserving defor- +mations play a key role for the semiclassical solution of +the eigenvalue equation (7). The group of all such de- +formations is obviously huge, so it is essential to identify +the subset of transformations that are likely to be im- +portant for low-energy physics. In fact, part of this work +has already been carried out, at least implicitly, in the +seminal series of papers [58–63], which we now use as a +basis for the definition of edge deformations. (A similar +motivation was put forward in [57].) +Label points on the plane by their polar coordinates +(r, ϕ), defined as usual by x + iy = r eiϕ. +Then, the +boundary of any isotropic QH droplet is located at some +fixed radius redge = O(ℓ +√ +N). What is the most general +area-preserving deformation that preserves this order of +magnitude? The answer is readily found by realizing that +the constraint of preserving redge = O(ℓ +√ +N) is equiva- +lent, at leading order in 1/N, to the condition that the de- +formation commutes with overall dilations r �→ const×r. +The most general deformation satisfying this criterion is +an edge deformation +�r2 +2 , ϕ +� +�→ +� +r2 +2f ′(ϕ), f(ϕ) +� +, +(23) +where f(ϕ) is an (orientation-preserving) deformation of +the circle, i.e. any smooth map satisfying f(ϕ + 2π) = + +7 +f(ϕ) + 2π and f ′(ϕ) > 0. The angle-dependent rescaling +of r on the right-hand side ensures that the map preserves +area, and reproduces the argument of the potential (10) +with λ = 2f ′. Note that the set of maps (23) is isomor- +phic to the group of diffeomorphisms of the circle, whose +central extension famously leads to the Virasoro algebra +encountered in CFT. Indeed, this motivates the state- +ment in [60, 61] that generators of maps (23) in the QH +effect produce conformal transformations of edge modes. +We stress that the subset of transformations (23) orig- +inates from an asymptotic analysis of the relevant or- +ders of magnitude. One can undoubtedly consider other +families of deformations, motivated by different consid- +erations, but those are irrelevant for our purposes. For +instance, the transformations r2 �→ r2 + α(ϕ) are crucial +for the effective low-energy description of QH droplets +[6, 10, 61], but they are subleading compared to (23) since +they deform the radius redge = O(ℓ +√ +N) by terms of order +O(1/N) instead of O(1). Conversely, one might consider +‘higher-spin transformations’ [58, 60, 61] that change the +radius in a dramatic way such as r2 �→ β(ϕ)r4[1+O(1/r)], +but these stretch QH droplets to an infinite extent in the +thermodynamic limit, which is why we discard them. +Let us provide a few examples of edge deformations +for future reference. First, (23) includes rigid rotations +around the origin given by f(ϕ) = ϕ + const. A richer +class is obtained by fixing some positive integer k and +considering all maps of the form +eikf(ϕ) = α eikϕ + β +¯β eikϕ + ¯α +(24) +where α, β are complex and satisfy |α|2 − |β|2 = 1. For +fixed k, such maps span a group locally isomorphic to +SL(2, R), always containing a subgroup of rigid rotations. +We will return to these deformations below, since they +can be seen as Fourier modes for circle diffeomorphisms. +In particular, setting α = cosh λ and β = sinh λ for some +real parameter λ turns the map (24) into an analogue of +a Lorentz boost with rapidity λ. In terms of the bulk +action (23), any deformation (24) turns a circle into a +‘flower’ with k petals: see Fig. 5 for k = 3. For k = 2, +this maps the circle on an ellipse [57], which will be useful +for anisotropic harmonic traps in Sec. IV E. +B. +Edge-deformed potentials +Given an isotropic potential V0(r2/2), how is it affected +by an edge deformation (23)? The answer is provided by +the anisotropic trap (10) with λ(ϕ) = 2f ′(ϕ): +V (r, ϕ) ≡ V0 +� +r2 +2f ′(ϕ) +� +. +(25) +In what follows, we exclusively consider this class of po- +tentials and refer to them as ‘edge-deformed traps’, for +the reasons stated above. Having fixed once and for all +some circle deformation f(ϕ), our goal is to solve the cor- +responding eigenvalue equation (7) in the classical limit +of high quantum numbers and small magnetic length. +We begin by listing the key classical data of the prob- +lem. The action-angle variables corresponding to (25) are +(ℓ2K, θ) = +� +r2� +(2f ′(ϕ)), f(ϕ) +� +with an inverse given by +(r2/2, ϕ) = +� +ℓ2K/(f −1)′(θ), f −1(θ) +� +, where f −1 denotes +the 1D inverse of f. Points at constant K are equipoten- +tials, i.e. level curves of (25), each of which is a set of +points such that +r2 +2f ′(ϕ) = ℓ2K +(26) +with constant K ≥ 0. +In Cartesian coordinates, this +is the set of points x = +� +2ℓ2Kf ′(ϕ) cos(ϕ), y += +� +2ℓ2Kf ′(ϕ) sin(ϕ) for ϕ ∈ [0, 2π]. +Equivalently, in +terms of the angle variable θ = f(ϕ) ∈ [0, 2π], the equipo- +tential is +x = +� +2ℓ2K +(f −1)′(θ) cos(f −1(θ)) ≡ F(K, θ), +y = +� +2ℓ2K +(f −1)′(θ) sin(f −1(θ)) ≡ G(K, θ), +(27) +where the notation (F, G) was introduced in Sec. III A. +Note that we will eventually focus on the regime where +K ≫ 1 is very large in such a way that the dimensionful +area 2πℓ2K be an O(1) quantity as ℓ → 0. +Moving just slightly away from the classical regime, we +have seen in Sec. III that the expansion of the operator +P V P involves a bicomplex potential function V(z, ¯w). In +the case of edge-deformed potentials (25), with the con- +ventions used there and above for complex coordinates, +one finds +V(z, ¯w) = V0 +� +ℓ2 +z ¯w +f ′� 1 +2i log[z/ ¯w] +� +� +. +(28) +Note that this only makes sense for z and w close to each +other, otherwise taking z → e2πiz affects the argument +of f ′ on the right-hand side. By contrast, when z and w +remain close, taking z → e2πiz also requires w → e2πiw, +and this time the angle +1 +2i log[z/ ¯w] is indeed invariant. +Finally, the expansion (22) also involves the complexi- +fied Laplacian of the potential, but only its real value will +be relevant at the order studied here. Let us therefore +express the Laplacian of (25) in polar coordinates: +∇2V = 1 +f ′ +� +2 − 1 +2 +f ′′′ +f ′ + f ′′2 +f ′2 +� +V ′ +0 +� +r2/2f ′� ++ r2 +f ′2 +� +1 + f ′′2 +4f ′2 +� +V ′′ +0 +� +r2/2f ′� +. +(29) +Here the prime means differentiation with respect to the +argument, namely ϕ for f(ϕ) and r2/2 for V0(r2/2). We +shall rely on (28) and (29) below, since they directly affect +the eigenvalue equation (7). + +8 +C. +Eigenvalue equation and energy +Having studied the potential (25), let us turn to the +quantum state meant to solve the eigenvalue equation +(7). As in Sec. III B, we begin by building a state (18) +that is perfectly localized on the equipotential, project +to the LLL using the operator (5), and obtain the wave +function (20) that now reads +ψm(z, ¯z) = e−|z|2/2 +� +dϕ f ′(ϕ) n(f(ϕ)) +× exp +� +imf(ϕ) − 1 +2mf ′(ϕ) + z +� +mf ′(ϕ) e−iϕ� +, +(30) +where we changed variables using θ = f(ϕ). It remains to +show that this solves the eigenvalue equation (7) for edge- +deformed traps (25) in the semiclassical regime, provided +the function n(θ) satisfies a suitable transport equation. +The latter is derived by expanding the energy (9) and +the potential (22) to get +0 = +� +dϕ f ′(ϕ) n(f(ϕ)) +× exp +� +imf(ϕ) − 1 +2mf ′(ϕ) + z +� +mf ′(ϕ) e−iϕ� +× +� +V +� +z, +� +mf ′(ϕ) e−iϕ� ++ ℓ2 +2 ∇2V − E0 +m − ℓ2E1 +m +� +(31) +where V(z, ¯w) is the bicomplex function (28) and the +equation holds up to neglected O(ℓ4) corrections. +At +leading order in the classical limit, the potential expan- +sion (22) boils down to ⟨z|P V P|w⟩ ∼ V(z, ¯z)δ2(z − w), +so (31) merely states that E0 +m = V0(ℓ2m) = V (γm). The +issue is to find the two remaining unknowns: the function +n(f(ϕ)) and the first-order energy correction E1 +m. +To determine these, the crucial step is to evaluate (31) +along the equipotential (26) labeled by K = m, i.e. for +z = +� +mf ′(α) eiα with α ∈ [0, 2π), where as usual we +assume m ≫ 1. Indeed, if (31) holds on a level curve, +then it holds for all z by holomorphicity. This is writ- +ten in more detail in App. C, where we show that the +integrand of (31) has a saddle point at ϕ = α, eventu- +ally resulting in a transport equation for the unknown +function n. Here we skip the computation and analyse +separately the real and imaginary parts of the transport +equation. We start with the real part, which will allow +us to deduce the LLL-projected energy spectrum. The +imaginary part is postponed to Sec. IV D, where we also +display the resulting nearly Gaussian wave functions. +Let Φ(ϕ) denote the phase of n(f(ϕ)) ≡ N(ϕ) eiΦ(ϕ). +Then the real part of the transport equation [see (C17)] +yields +Φ′(ϕ) = E1 +m +Ωm +f ′(ϕ) − 1 +2 +� Γm +Ωm ++ 1 +�� +1 + f ′′(ϕ)2 +4f ′(ϕ)2 +� +− 1 +2 + ∂ϕ +� f ′′(ϕ) +8f ′(ϕ) +� ++ 1 +2 +∂ϕ[f ′′(ϕ)/2f ′(ϕ)] +1 + f ′′(ϕ)2/4f ′(ϕ)2 , +(32) +where E1 +m is the first order correction to the energy (9) +and we introduced the parameters +Ωm ≡ V ′ +0(ℓ2m) > 0, +Γm ≡ ℓ2m V ′′ +0 (ℓ2m). +(33) +In many-body droplets with N electrons, these will re- +spectively measure the Fermi velocity and the curvature +of the potential at the Fermi surface when m = N. Note +that all terms in (32) are total derivatives save for the +factor 1 + [f ′′/2f ′]2, so the solution is +Φ(ϕ) = E1 +m +Ωm +f(ϕ) − 1 +2 +� Γm +Ωm ++ 1 +� � ϕ +0 +dθ +� +1 + f ′′(θ)2 +4f ′(θ)2 +� +− ϕ +2 + f ′′(ϕ) +8f ′(ϕ) + 1 +2 arctan +� f ′′(ϕ) +2f ′(ϕ) +� ++ const. +(34) +This turns out to imply a quantization condition for en- +ergy. Indeed, when we initially introduced the function +n(θ) in (18), we mentioned that it must have a vanishing +winding number along the equipotential, so that all the +winding of the wave function is contained in the expo- +nential factor eimθ. The phase Φ(ϕ) must therefore be +strictly 2π-periodic, i.e. Φ(2π) = Φ(0). Using (34), this +fixes the first quantum correction of the energy (9): +E1 +m +Ωm += 1 +2 + +� Γm +Ωm ++ 1 +� � dϕ +4π +� +1 + f ′′(ϕ)2 +4f ′(ϕ)2 +� +. +(35) +The latter generally generally depends on m through Γm +and Ωm in (33). +A simplification only occurs in ‘har- +monic’ setups where Γm = 0 and the right-hand side of +(35) is an f-dependent constant, for all m [72]. In any +case, the full mth energy (9) in the semiclassical limit can +be written as +Em ∼ V0 +� +ℓ2m +� ++ ℓ2 +2 +� +Ωm + +� +Γm + Ωm +� � dϕ +2π +� +1 + f ′′(ϕ)2 +4f ′(ϕ)2 +�� +, +(36) +reproducing the result announced in (12)–(13) with λ = +2f ′ and generalizing the isotropic value obtained for +f ′ = 1 [see (A3)]. The leading-order Bohr-Sommerfeld +quantization condition (8) is manifestly satisfied, while +the first quantum correction can be expressed in terms +of a Maslov-like shift and an integral of the Laplacian, +confirming the general result in [55]: +Em = V0 +� +ℓ2� +m + 1 +2 +�� ++ ℓ2 +4 +� dϕ +2π f ′(ϕ)∇2V +�� +r2=2ℓ2(m+1/2)f ′(ϕ) + O(ℓ4). +(37) +(In the language of [55], our ‘Maslov-like’ term actually +stems from an integral of the curvature of γm.) It now +remains to write the corresponding wave functions. +D. +Gaussian wave functions +As above, write n(f(ϕ)) = N(ϕ) eiΦ(ϕ) for the un- +known function of the WKB ansatz, with a norm N(ϕ) = + +9 +|n(f(ϕ))|. +Then the imaginary part of the transport +equation [see (C18)] can be recast into +N ′(ϕ) +N(ϕ) = 1 +4∂ϕ log +� +1 +f ′(ϕ) +� +1 + f ′′(ϕ)2 +4f ′(ϕ)2 +�� +, +(38) +which remarkably has the form of an overall logarithmic +derivative. The general solution is thus +��n +� +f(ϕ) +��� = N0 +� +1 +f ′(ϕ) +� +1 + f ′′(ϕ)2 +4f ′(ϕ)2 +��1/4 +, +(39) +where the normalization N0 will soon be fixed. Note the +exponent 1/4, typical of WKB approximations [56]. +We can now use (39) to evaluate approximate eigen- +functions (30) near their maximum, i.e. close to the +equipotential (C2). To see this, zoom in on the equipo- +tential by writing +z ≡ +�√m + a +�� +f ′(α) eiα +(40) +for m ≫ 1 and some finite a. The integral (30) then has +a unique saddle point at ϕ = α−iδ1/√m+O(1/m), with +δ1 = a +� +1 − i f ′′(α) +2f ′(α) +�−1 +. The saddle-point approximation +of the wave function (30) thus yields +ψm(z, ¯z) ∼ +1 +√ +2πℓ2 +1 +(2πm)1/4 eimf(α)+iΦ(α) +× +1 +� +σ(α) +exp +� +�− f ′(α) a2 +1 − i f ′′(α) +2f ′(α) +� +� , +(41) +where we used Eqs. (34)–(39) for the phase and norm of +n(f(α)), fixed the integration constant in (39) to N0 = +1 +2πℓ( m +2π)1/4, and introduced the ubiquitous width +σ(ϕ)2 ≡ +1 +f ′(ϕ) +� +1 + f ′′(ϕ)2 +4f ′(ϕ)2 +� +. +(42) +This is the angle-dependent variance of the probability +density of (41), written in (12) in terms of λ = 2f ′. In- +deed, one has +|ψm(z, ¯z)|2 ∼ +1 +2πℓ2 +e−2a2/σ2(α) +√ +2πm σ(α), +(43) +which +satisfies +the +desired +normalization +condition +� +d2x |ψ|2 = 1. Note that Eq. (41) reproduces the wave +function stated in (11), once more with λ = 2f ′ and now +involving the phase +Θm(x) = mf(ϕ) + Φ(ϕ) − +a2f ′′(ϕ) +2f ′(ϕ)σ(ϕ)2 . +(44) +The Gaussian behavior of LLL-projected is thus mani- +fest, as anticipated at the end of Sec. III B for the gen- +eral WKB ansatz (20). It is illustrated in Fig. 3 for two +choices of the confining potential (25). Finally, (41) gen- +eralizes the behavior of isotropic states (3) [see (A1)], in- +cluding the O(1/√m) contribution that we did not state +here but that can be computed by incorporating the next +order term δ2/m for the saddle point and repeating the +analysis [73]. +Out[� ]= +0 +0.2 +0.4 +0.6 +0 +1 +|ψm|2 +Figure 3. The density of a wave function (41) at m = 30 in an +edge-deformed trap (25). The Gaussian behavior is manifest, +as is the angle-dependent ‘roller coaster’ predicted by (43). +Left: Elliptic potential given by (25) with f of the form (24) +and k = 2. Right: Same edge-deformed trap as in Figs. 1–2. +E. +Comparison with elliptic wave functions +To conclude this section, we now focus on the ‘flower’ +deformations (24) and show that the resulting transport +equation is integrable: both the phase (34) and the norm +(39) can be expressed in terms of elementary functions. +These results are valuable in themselves since flower maps +are the simplest edge deformations—they are analogues +of Fourier modes for circle diffeomorphisms—, but also +because their special case k = 2 reproduces known wave +functions in elliptic harmonic traps [68, 74], providing an +important benchmark for our WKB approach. +Consider first the deformation (24) with α = cosh λ +and β = sinh λ for an arbitrary integer k and a real pa- +rameter λ. Then the energy quantization condition (35) +can be integrated exactly, yielding +E1 +m +Ωm += 1 +2 + 1 +2 +� +1 + Γm +Ωm +�� +1 + k2 +4 +� +cosh(2λ) − 1 +�� +. (45) +As for the solution of the transport equation, consisting +of the phase (34) and norm (39), it is found to be +n(θ) = N0 e +i +8 +Γm +Ωm k sin(kθ) sinh(2λ)e +i θ +2 +� +( +Γm +Ωm +1) +� +1− k2 +4 +� ++1− k +2 +� +× +�cosh λ + eikθ sinh λ +eikθ cosh λ + sinh λ +� k2−4 +8k (1+ Γm +Ωm )− 1 +2k + 1 +4 +× +� +−4eikθ + (e2ikθ − 1)k sinh(2λ) +eikθ cosh λ + sinh λ +(46) +up to an overall constant phase. +Note that (46) depends in a non-trivial way on the po- +tential derivatives (33), with some simplification in the +‘harmonic’ regime Γm = 0. Let us therefore apply Eqs. +(45)–(46) to the case of an elliptic potential k = 2 with +constant stiffness Ωm = Ω > 0 (hence Γm = 0). The cor- +responding edge deformation (23) maps the isotropic har- +monic potential V0(x) = Ω r2/2 on its anisotropic cousin, +V (x) = Ω e−2λx2 + e2λy2 +2 cosh(2λ) +, +(47) + +10 +whose equipotentials are ellipses rather than circles. The +energy correction (45) then becomes E1 +m = +1 +2Ω[1 + +cosh(2λ)] and the solution (46) of the transport equation +can be written as +n(θ) = N0 +� +cosh λ − e2iθ sinh λ. +(48) +It is straightforward to use this data to obtain the elliptic +version of the normalized Gaussian wave function (41): +ψm(z, ¯z) ∼ +�2π +m +�1/4 +1 +2πℓ +eimθ +√ +cosh λ − e−2iθ sinh λ +× exp +� +−e2iθ + tanh(λ) +e2iθ − tanh(λ)a2 +� +. +(49) +Crucially, the latter coincides with the large m approxi- +mation of the exact LLL-projected eigenstates of the har- +monic potential (47), as can be verified thanks to known +asymptotic formulas for Hermite polynomials [68]. This +is actually true even at subleading order in 1/√m, which +we omit here for brevity. +V. +MANY-BODY OBSERVABLES +This section finally applies the results of Secs. III–IV +to fully-fledged QH droplets consisting of a large num- +ber N ≫ 1 of electrons. +Specifically, we exploit our +insights on near-Gaussian single-particle wave functions +(41) to compute many-body observables and read off the +universal shape-dependent effects implied by the width +(42). The density is treated first: it equals +1 +2πℓ2 in the +bulk, then drops to zero as an error function at the edge +redge = ℓ +� +2Nf ′(ϕ). We then turn to the current and +show that it is localized as a Gaussian on the edge, to +which it is tangent. +Third, correlations near the edge +are found to display the usual power-law behavior of free +fermions, dressed by radial Gaussian factors. +This re- +duces to known expressions in isotropic traps [59] and in +the case of flower deformations (24) with k = 2, where +one recovers the elliptic results of [68, 74]. Finally, the +radial behavior of correlations is shown to be consistent +with the effective low-energy field theory of edge modes: +we derive it microscopically and obtain a chiral CFT in +terms of the angle variable on the boundary. +A. +Density +Consider a QH droplet of N ≫ 1 non-interacting 2D +electrons subject to the Hamiltonian (1), with a very +strong magnetic field B = dA and a weak edge-deformed +potential (25). The ground state of this many-body sys- +tem is a Slater determinant of the wave functions ψm for +occupied states m = 0, 1, . . . , N −1, where we recall that +m is a quantized action variable generalizing angular mo- +mentum (see the red dots in Fig. 4). Each ψm yields a +m +Energy +◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦•••• +N +Fermi energy +••••••••◦◦◦◦◦◦· · · +◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦◦· · · +2Λ +Figure 4. The one-body spectrum (37), where the states that +are occupied in the many-body ground state are highlighted in +red and those that contribute to the low-energy Hamiltonian +(64) are filled (black for ‘particles’, red for ‘holes’). The cutoff +Λ is large but much smaller than N in the sense that Λ = O(1) +in the thermodynamic limit. +single-particle probability density |ψm(x)|2, the sum of +which gives the many-body density +ρ(x) = +N−1 +� +m=0 +|ψm(x)|2. +(50) +While WKB theory does not give access to the form of ψm +at low m, large values of m should be correctly captured +by the analysis of Sec. IV, in which case the one-body +density is approximately Gaussian and given by (43). We +now exploit this Gaussian form to evaluate the many- +body density, both in the bulk and close to the edge. +(Some technical details are highlighted along the way, as +the same method will later allow us to study the many- +body current and correlations.) +The key point is that each wave function (43) is local- +ized on an equipotential of V (x) with area 2πℓ2m, so the +density close to some equipotential |z| = const × +� +f ′(ϕ) +only receives sizeable contributions from wave functions +whose quantum number is close to |z|2/f ′(ϕ). Accord- +ingly, the bulk density for 1 ≪ |z| ≪ +√ +N is obtained by +letting the upper summation bound of (50) go to infinity +and writing the approximate density as +ρ(x) ∼ +1 +2πℓ2 +∞ +� +m=m0 +e +− +2 +σ2(ϕ) +� +|z| +√ +f′(ϕ) −√m +�2 +√ +2πm σ(ϕ) +, +(51) +where the lower summation bound m0 ≫ 1 is irrele- +vant as long as it is much smaller than |z|2, and σ(ϕ) is +the angle-dependent width (42). At large |z|, the Euler- +Maclaurin formula allows us to approximate the sum over +m by a (Gaussian) integral over √m. This yields the ex- +pected density +ρ(x) ∼ +1 +2πℓ2 +(52) +in the bulk of a QH droplet with filling fraction ν = 1. +An analogous argument can be carried out close to +the droplet’s edge, with one key difference: the upper +summation bound of (50) is now crucial. Thus, letting +|z| = +�√ +N + a +� +f ′(ϕ) with finite a in the large N limit + +11 +and using once more the approximate Gaussian behavior +(43), the density (50) near the edge behaves as +ρ(x) ∼ +1 +2πℓ2 +∞ +� +k=1 +e +− +2 +σ2(ϕ) +� +a+ +k +2 +√ +N +�2 +√ +2πN σ(ϕ) +, +(53) +where we changed variables as m ≡ N − k with k = +O( +√ +N) at large N and only kept track of leading-order +terms. For N ≫ 1, the sum over k can once more be +converted into an integral, now over k/2 +√ +N. This yields +the asymptotic behavior +ρ(r, ϕ) ∼ +1 +4πℓ2 erfc +� +1 +σ(ϕ) +r − ℓ +� +2Nf ′(ϕ) +ℓ +� +f ′(ϕ) +� +(54) +where erfc denotes the complementary error function and +the width (42) is inherited from that of LLL wave func- +tions. This is a remarkably explicit result, announced in +(14) with λ = 2f ′. It confirms that the density is roughly +constant and equal to (52) in the bulk, then drops to zero +within a distance of the order of the magnetic length (2) +around the edge at r = ℓ +� +2N f ′(ϕ). See Figs. 2(a)–5(a). +We stress that, in contrast to wave functions, the den- +sity (54) only depends on the potential near the edge of +the droplet: bulk deformations of the potential do not af- +fect the quantized bulk density (52) in the limit of strong +magnetic fields. In this sense, (54) is a universal formula +for the density of any QH droplet of LLL states whose +edge traces an equipotential of the form r2 = 2ℓ2N f ′(ϕ). +It would be instructive to probe this local density in mod- +ern quantum simulators [21–24]. +Note that the leading-order formula (54) receives a +number of subleading corrections that can be systemati- +cally computed in our formalism; these are omitted here +for brevity. A related comment is that the bulk value +density (52) is only valid at extremely strong magnetic +fields, which stems from the simplification provided by +the LLL projection. The actual density profile depends +on the gradient of the potential, but this involves higher +Landau levels that are beyond our scope [28, 29]. +B. +Current +The current of a droplet of N ≫ 1 electrons can simi- +larly be evaluated as a sum over single-particle currents. +To this end, recall that the gauge-invariant one-body +probability current of a charged wave function ψ with +mass M is a one-form ℏ j/M given by +j = 1 +2i +� +ψ∗dψ − ψdψ∗ − 2iq +ℏA|ψ|2� +, +(55) +where the first term is only sensitive to the phase of ψ +and A = 1 +2Br2 dϕ = ℏ +q |z|2 dϕ in symmetric gauge. Thus, +the many-body current of a Slater determinant of the +occupied states ψm with m = 0, . . . , N − 1 is +J = +N−1 +� +m=0 +jm, +(56) +where jm is the single-particle current (55) of each ψm. +As before, the WKB approximation does not give ac- +cess to wave functions for small m, but this is unim- +portant close to the edge. In that regime, we have al- +ready gathered all ingredients needed to evaluate the +currents (55) up to small quantum corrections: the one- +body density is given by (43), while the derivative of +the phase is contained in (41) and the phase transport +equation (32). +In practice, the WKB phase Φ turns +out to be negligible at leading order, and the only rel- +evant parts of the phase are those already visible in (41): +the (fast) phase eimf(ϕ) together with the contribution +from A = +� +ℏ|z|2/q +� +dϕ eventually gives rise to the lead- +ing ϕ component of the current, while the (slow) phase +e−i[f ′′(ϕ)/2f ′(ϕ)]a2/σ2 yields its radial component that is +non-zero whenever f ′′(ϕ) ̸= 0. +Starting from these facts, it is straightforward to adapt +the method of Sec. V A to the many-body current (56). +Writing |z| = +�√ +N +a +�� +f ′(ϕ), the sum over m ≡ N −k +becomes an integral over k/(2 +√ +N) = O(1), which yields +the leading order result quoted in (15) with λ = 2f ′: +J(r, ϕ) ∼ − +e +− +2a2 +σ(ϕ)2 +(2πℓ2)3/2σ(ϕ) +ℓ +� +2Nf ′(ϕ) dϕ + f ′′(ϕ) +2f ′(ϕ) dr +� +2f ′(ϕ) +(57) +where a = +� +r−ℓ +� +2Nf ′(ϕ) +�� +ℓ +� +2f ′(ϕ) and σ(ϕ) is given +by (42). Both components in (57) receive subleading cor- +rections that are omitted here. In particular, there is an +O(1) term in Jϕ that is non-zero on the edge, even in the +isotropic case f ′ = 1. The computation of that term re- +quires the O(1/√m) correction that was omitted in (43). +Using the metric ds2 = dr2 + r2dϕ2, one verifies that +the one-form ℓ√2N f ′ dϕ+(f ′′/2f ′) dr in (57) is the dual +of a vector tangent to the equipotential at the droplet’s +edge [75]. Moreover, the norm squared of (57) is +∥J(r, ϕ)∥2 ∼ +1 +2(2πℓ2)3 exp +� +−2 +� +r − ℓ +� +2Nf ′(ϕ) +�2 +ℓ2σ(ϕ)2f ′(ϕ) +� +, +(58) +showing that the current has a constant maximum along +the edge but a varying width: see Fig. 5. +Similarly to the density, it is important to remember +that the LLL projection misses some important physics. +Indeed, the actual bulk current is the symplectic gra- +dient of the confining potential multiplied by the Hall +conductance [28, 29, 41, 42]. +No such effect occurs in +(57) because it requires higher Landau levels, which are +beyond our scope. +C. +Correlations +The methods that we have applied to density and cur- +rent can also be used to compute electronic correlations +near the edge. Indeed, consider as before an anisotropic +droplet whose occupied one-body states have quantum + +12 +��� +��� +��� +x +x +x +y +y +y +√ +N +− +√ +N +− +√ +N +√ +N +(a) +(b) +(c) +x +x +x +y +y +y +ℓ +√ +2N +−ℓ +√ +2N +−ℓ +√ +2N +ℓ +√ +2N +Figure 5. (a) The density (54) for N = 100 electrons and a ‘flower’ edge deformation (24) of order k = 3. The constancy of +density in the bulk and its sharp decay at the boundary are manifest. (b) The current’s norm (58) for the same flower-shaped +droplet. The localization on the edge equipotential (black line) is clearly visible, as is the position-dependent width of the +Gaussian jump. (c) The correlation function C(x1, x2) (61) for the same flower-shaped droplet, seen as a function of x2 when +x1 = (ℓ +� +Nλ(0), 0) is fixed close to the edge of the droplet. +numbers m = 0, 1, . . . , N − 1. Then the correlation func- +tion between the points x1 and x2 is +C(x1, x2) = +N−1 +� +m=0 +ψ∗ +m(x1) ψm(x2), +(59) +which reduces to the density (50) when x1 = x2. +As +before, we rename m ≡ N − k and let the complex coor- +dinates z, w corresponding to x1, x2 be such that +z = +�√ +N + a +�� +f ′(ϕ1) eiϕ1, +w = +�√ +N + b +�� +f ′(ϕ2) eiϕ2, +(60) +where a, b are finite at large N and ϕ1, ϕ2 are the po- +lar angles of x1, x2. +One can then plug the Gaussian +wave functions (41) into (59), this time assuming k fi- +nite, and performing the sum over k. The gradient ex- +pansion of the potential implies that the ratio Γm/Ωm ∼ +ΓN/ΩN + O(ℓ2) is nearly constant in this regime, so Eq. +(59) becomes a geometric sum over k that reproduces the +result stated in (16) with λ = 2f ′: +C(x1, x2) ∼ +eiΘN(x1,x2) +(2π)3/2ℓ2√ +N +1 +� +σ(ϕ1)σ(ϕ2) +× +i e +− +a2 +σ(ϕ1)2 − +b2 +σ(ϕ2)2 +2 sin +� +[f(ϕ1) − f(ϕ2)]/2 +�, +(61) +where σ(ϕ) is the angle-dependent width (42). The over- +all phase ΘN(x1, x2) = ΘN(x2) − ΘN(x1) is given by +(44), and involves in particular the WKB phase (34). +Some features of (61) are worth emphasizing. First, +note +the +Gaussian +jump +of +power-law +correlations +near the edge, involving a free fermion correlation ∝ +sin([f(ϕ1)−f(ϕ2)]/2)−1 expressed in terms of f(ϕ1) and +f(ϕ2); this is the standard static diagnostic of the pres- +ence of edge modes [10, 59, 76]. A second striking aspect +is the apparent lack of translation-invariance along the +edge, caused not only by the argument f(ϕ1) − f(ϕ2) = +� ϕ1 +ϕ2 dθ f ′(θ) but also by the widths σ(ϕ1) and σ(ϕ2). In +particular, the product σ(ϕ1)−1/2σ(ϕ2)−1/2 is reminis- +cent of prefactors picked up by primary fields in CFT +under conformal maps. +Since the boundary correlator (61) holds in any edge- +deformed trap (25), it also applies to special cases of in- +terest such as the anisotropic harmonic potential (47). +The corresponding correlations were actually computed +long ago in [68], and they coincide with our result (61) +upon using the map (24) with k = 2, α = cosh λ, +β = sinh λ. In fact, this specific setup is also well known +in the context of the Coulomb gas, since edge correla- +tions can then be related by a conformal map to the +standard Euclidean correlation function (z1 − z2)−1 of a +free fermion CFT [77]. +Finally, +it is a simple matter to include time- +dependence in the correlator (61). Indeed, the occupied +one-particle states in (59) have definite energies Em given +by (37) at large m. This spectrum is approximately linear +close to the Fermi energy: changing variables according +to m = N + k with k finite at large N, one has +EN+k − EN ∼ ℏω k +(62) +where ω ≡ ℓ2ΩN/ℏ is the angular Fermi velocity given +by the derivative (33) at m = N. +Note that ω re- +ceives a number of subleading quantum corrections in- +volving e.g. the curvature ΓN in (33) [78–80]; we ne- +glect those. +In the linear regime (62), one can repeat +the asymptotic computation of the correlator and find + +- +1 +1 +1 +1 +- +4 +r +- +1 +1 +1 +1 +1 +1 +- +1 +- +1 +1 +- +1 +1 +1 +113 +once more a result of the form (61), now with a time- +dependent overall phase and a time-dependent denom- +inator 2 sin +� +[f(ϕ1) − f(ϕ2) − ω(t1 − t2)]/2 +� +. +This ex- +hibits the standard ballistic propagation of correlations +in a CFT, which we confirm below by deriving the effec- +tive low-energy dynamics of our droplet. +D. +Edge modes +The effective low-energy description of anisotropic QH +droplets can be derived similarly to the isotropic case +[59] inspired by Luttinger-liquid theory [81]. +This has +the advantage of circumventing topological field theory, +at the cost of failing to apply in fractional QH states [6– +10]. We now provide such a first-principles calculation, +eventually concluding that edge modes span a free chiral +CFT expressed in terms of the angle coordinate f(ϕ) +along the boundary. +Aside from its intrinsic interest, +this computation provides an independent check of the +validity of the correlation (61). +Our starting point is the second-quantized Hamilto- +nian in a Fock space of free fermions, +H = +� +m,n +(Em,n − µ)a† +m,nam,n. +(63) +Here the sum runs over eigenstates of the one-body +Hamiltonian (1), labeled by the Landau level n ∈ N +and the ‘action variable’ quantum number m ∈ N within +each level. Their energies are denoted Em,n, and µ is +some chemical potential to be fixed below. +For each +pair (m, n), the operator a† +m,n creates the corresponding +eigenstate. The exact energy spectrum is unknown, but +this is not an issue since low-energy excitations all be- +long to the LLL, with an approximately linear dispersion +(62) within a window [−Λ, Λ] around the Fermi momen- +tum. Then choosing the chemical potential of (63) as +µ = const − 1 +2ℏω, the low-energy approximation of the +many-body Hamiltonian (63) becomes +H ∼ +� +k∈[−Λ,Λ] +ℏω (k + 1/2) a† +N+kaN+k, +(64) +where the sum is over all integers k in the specified win- +dow. Since aN+k are fermionic Fock space operators, the +Λ → ∞ limit of (64) yields a gapless Hamiltonian for free +fermions written in Fourier modes, and the chemical po- +tential enforces antiperiodic (Neveu-Schwarz) boundary +conditions. It now only remains to relate the edge CFT +to bulk wave functions. +To this end, note that the Fock space operators in (64) +create states in the LLL that are given by the Gaussian +wave function (41). One can therefore express them in +terms of 2D creation operators c†(x): +a† +N+k ∼ +� f ′(ϕ) dϕ +√ +2π +ei(k+1/2)f(ϕ) Ψ†(f(ϕ)), +(65) +where the 1D fermionic field Ψ is defined as a radial in- +tegral involving the wave function (41), +Ψ†(f(ϕ)) ≡ ei(N−1/2)f(ϕ) eiΦ(ϕ) +f ′(ϕ) +� +σ(ϕ) +× +� ∞ +0 +r dr +ℓ(2πN)1/4 c†(x) e +− +1 +σ2(ϕ) +� +r +√ +2ℓ2f′(ϕ) − +√ +N +�2 +(66) +with σ the width (42) and Φ the WKB phase (34). Note +that in writing (66) we assumed that the momentum in- +dex k and the cutoff Λ are of order O(1) in the thermo- +dynamic limit. It is then clear that the operator Ψ†(ϕ) +creates an electron at the position ϕ on the edge. +From this point onward, the derivation of the low- +energy effective field theory is essentially done. Indeed, +the Hamiltonian (64) expressed in terms of the edge field +(66) reads +H = ℏω +� +dθ Ψ†(θ) (−i∂θ)Ψ(θ). +(67) +Furthermore, the normalization of the field Ψ defined by +(66) is canonical in angle variables: using the standard +anticommutator {c(x1), c†(x2)} = δ(2)(x1 −x2), one sim- +ilarly finds {Ψ(f(ϕ1)), Ψ†(f(ϕ2))} = δ(f(ϕ1) − f(ϕ2)) in +terms of the Dirac delta function on a circle. This de- +termines the kinetic term of the action functional for Ψ, +which is thus a canonical fermionic expression ∼ Ψ†∂tΨ. +It can be combined with the Hamiltonian (64) to write +down the fermionic action functional of edge modes, +S[Ψ, Ψ†] = ℏ +� +dt dθ iΨ†(θ) +� +∂t + ω∂θ +� +Ψ(θ), +(68) +where we recall that the angular Fermi velocity is ω = +ℓ2ΩN/ℏ, given by (33). This is manifestly a (1+1)D free +chiral CFT in terms of the angle variable θ = f(ϕ). +The low-energy effective theory (68) is universal: for +any trapping potential, edge modes are described by a +chiral fermionic CFT expressed in terms of the angle co- +ordinate of the trap at the boundary, as could have been +guessed from the dynamics of electronic guiding centers +induced by the potential V in the non-commutative plane +(6). In the present case, the angle coordinate was just +θ = f(ϕ); more general cases involve more complicated +action-angle variables. We stress that the angle variable +generally has nothing to do with other obvious position +coordinates, such as the polar angle ϕ or the arc length +s(ϕ) = ℓ +√ +2N +� ϕ +0 +dα +� +f ′(α) + f ′′(α)2 +4f ′(α) . +(69) +Any such ‘wrong’ coordinate makes the apparent Fermi +velocity of edge modes position-dependent. This is rem- +iniscent of inhomogeneous CFTs whose light-cones are +curved owing to the presence of a non-zero space-time +curvature [82–89]. +However, one should keep in mind +that our edge modes sense a flat metric ω2dt2 − dθ2 = + +14 +ω2dt2 − f ′(ϕ)2dϕ2 whose light-cones are straight lines in +terms of the angle variable θ = f(ϕ). +Let us conclude this section by showing that the action +(68) is consistent with the seemingly complicated corre- +lator (61). We start from the definition (66) to write the +1D correlation function ⟨Ψ†(θ1)Ψ(θ2)⟩ as a double radial +integral of the 2D quantity ⟨c†(x1)c(x2)⟩. Now using the +asymptotic relation (61), one finds that all normaliza- +tions simplify and the correlator of the edge field boils +down to +⟨Ψ†(θ1)Ψ(θ2)⟩ = 1 +2π +i +2 sin +� +[θ1 − θ2]/2 +�. +(70) +The same result would have been obtained directly from +the low-energy action (68): it is a correlation function +of free gapless fermions written in terms of the angle +coordinates θ1 = f(ϕ1) and θ2 = f(ϕ2). As a bonus, +time-dependent correlations automatically satisfy the be- +haviour ∝ sin +� +[θ1 − θ2 − ω(t1 − t2)]/2 +�−1 stated at the +end of Sec. V C. +VI. +CONCLUSION AND OUTLOOK +This work was devoted to a detailed study of meso- +scopic droplets of non-interacting planar electrons placed +in a perpendicular magnetic field and confined by star- +shaped anisotropic traps with scale-invariant level curves. +In particular, we provided explicit formulas for the corre- +sponding wave functions and energy spectrum, allowing +us to compute the many-body density, current, and cor- +relations of an entire droplet. The resulting low-energy +edge modes were eventually shown to behave as a chiral +CFT in terms of the angle variable along the boundary. +This was all achieved in a regime of high magnetic fields, +ultimately equivalent to a semiclassical limit. +These results pave the way for a number of applications +and follow-ups. Indeed, recent advances in quantum sim- +ulation suggest the tantalizing possibility of probing lo- +cal properties of QH-like droplets in the lab [21–24], both +for static ground states and their dynamical edge excita- +tions. The density (54) or the current (57) then predict +observable shape-dependent widths, while the low-energy +theory (68) predicts the ballistic propagation of local +boundary disturbances with a position-dependent veloc- +ity. More generally, the geometry of the QH effect [33– +39] could soon become relevant for experiments involv- +ing ultracold atoms or photonics. Our framework pro- +vides a bridge between this field of mathematical physics +and concrete observables in mesoscopic quantum physics. +Verifying the predictions put forward here, through lin- +ear response experiments or direct imaging, would be a +fascinating example of many-body quantum mechanics +at work. +Turning now to theory, the link between our formal- +ism and QH symmetries deserves further study: following +the series of works [58–63], one can think of edge defor- +mations as approximately unitary operators acting on +many-body QH states. It is then natural to wonder how +these operators get composed together; do they span a +Virasoro group as in CFT? If yes, how to derive the cen- +tral charge c = 1 in terms of microscopic wave functions? +More broadly, what are the operators implementing area- +preserving deformations in the sense of the WKB ansatz +(20)? One expects these operators to provide a finite (ex- +ponentiated) form of the operators studied in [58, 60, 61], +with non-commutative composition laws consistent with +the geometry (6) of LLL physics. +Note that the discussion above was mostly focused on +leading-order properties. For instance, one may wonder +what are irrelevant corrections to the edge field theory +(68), especially following the recent numerical observa- +tion [79] that density waves on the edge satisfy a non- +linear Korteweg-de Vries equation at late times. +This +regime is presumably described by small droplet defor- +mations of the form r2 �→ r2 + α(ϕ), spanning a U(1) +Kac-Moody algebra whose level is sensitive to the fill- +ing fraction [58, 60, 61]. The resulting non-linear edge +waves would then be described by an evolution equa- +tion in an infinite-dimensional group manifold. This per- +spective is standard in hydrodynamics [90–92], but it +has only recently come to be appreciated in condensed +matter physics [93]. The geometric study initiated here +provides a basis for considerations of this kind in the +QH effect, including the possibility of inhomogeneous +(position-dependent) corrections in anisotropic traps. +Another obvious extension of this work is the fractional +QH regime. In that context, no single-particle descrip- +tion is available, but many-body predictions such as the +edge density (54) or the current (57) conceivably display +universal geometric features that would remain true in +interacting many-body ground states [42]. It would be +thrilling to derive such predictions from the family of +edge transformations studied here, either from a micro- +scopic analysis of the Laughlin wave function, or thanks +to the reformulation of fractional QH states as CFT cor- +relation functions [94]. +ACKNOWLEDGMENTS +We are grateful to Laurent Charles for illuminating +discussions on semiclassical methods in Kählerian geo- +metric quantization. B.O. also thanks Mathieu Beauvil- +lain, Nathan Goldman, and Marios Petropoulos for col- +laboration on closely related subjects. +Finally, we ac- +knowledge useful and motivating interactions with Jean +Dalibard, Benoit Douçot, Jean-Noël Fuchs, Marc Geiller, +Gian Michele Graf, Semyon Klevtsov, Titus Neupert +and Nicolas Regnault. The work of B.O. is supported +by the European Union’s Horizon 2020 research and in- +novation programme under the Marie Skłodowska-Curie +grant agreement No. 846244. B.L. acknowledges fund- +ing from the European Research Council (ERC) under +the European Union’s Horizon 2020 research and inno- +vation program (Grant No. ERC-StG-Neupert-757867- + +15 +PARATOP). P.M. gratefully acknowledges financial sup- +port from the Wenner-Gren Foundations (No. WGF2019- +0061). B.E. is supported by the ANR grant TopO No. +ANR-17-CE30-0013-01. +Appendix A: Isotropic droplets +Most of this work is concerned with anisotropic prop- +erties, so isotropic results provide a useful benchmark. +They are simpler than their anisotropic counterparts and +mostly well-known in the literature, so their properties +are concisely summarized here. We begin by recalling el- +ementary aspects of the one-body energy spectrum based +on the exact wave functions (3), then turn to many-body +observables. +1. +One-body spectrum +Consider a spin-polarized 2D electron governed by the +Landau Hamiltonian (1) with an isotropic confining po- +tential V (x) = V0(r2/2). At very strong magnetic fields, +the resulting one-body spectrum is well approximated +by the solution of the LLL-projected eigenvalue equa- +tion (7). As the potential is isotropic, it commutes with +angular momentum, so the eigenstates of P V P are wave +functions (3) with definite angular momentum. Note that +these confirm the general near-Gaussian behavior found +in (43): letting |z| = √m+a with finite a, one finds that +(3) behaves at large m as +φm(x) = +eimϕ +√ +2πℓ2 +1 +(2πm)1/4 e−a2 � +1 + O(1/√m) +� +. (A1) +The energy Em of each state (3) is readily found by com- +puting the wave function ⟨z, ¯z|PV0(r2/2)P|φm⟩, which +yields the exact eigenvalue +Em = ⟨φm|V |φm⟩ = 1 +m! +� ∞ +0 +dt tm e−t V0(ℓ2t) +(A2) +in terms of the integration variable t ≡ |z|2. Note in pass- +ing that this is the value one would find from first-order +perturbation theory of the full Landau Hamiltonian (1): +by construction, LLL-projected physics is only sensitive +to first-order effects of the potential, while higher-orders +ultimately involve higher Landau levels. +Now fix an index m ≥ 0. What is the corresponding +equipotential in the sense of (9)? To answer this in the +classical limit, we let m → ∞ while fixing the value of +ℓ2m = O(1), and evaluate the integral (A2) thanks to a +saddle-point approximation. The result is +Em = V0(ℓ2m) + ℓ2Ωm + ℓ2 +2 Γm + O(ℓ4), +(A3) +where Ωm and Γm were defined in (33). This is consistent +with Eqs. (13) and (36) with λ = 2f ′ = 2. +2. +Many-body aspects +The sequence followed here is the same as in Sec. V: we +start with the density, then consider the current and cor- +relations close to the edge. In all cases, the edge asymp- +totics reproduce the formulas of Sec. V for the simplest +case where f ′(ϕ) = 1. +Density. Let N ≫ 1 non-interacting planar electrons +be subjected to the Hamiltonian (1), with a very strong +magnetic field B = dA and a weak isotropic poten- +tial V (x) = V0(r2/2). +The ground state wave func- +tion of this many-body system is a Slater determinant of +the occupied single-particle eigenstates φ0, φ1, . . . , φN−1 +given by (3), each of which has a one-body density +|φm(x)|2. The resulting many-body density is thus (50), +which can be computed in closed form in the very spe- +cial case of states (3) with definite angular momentum: +it is a normalized incomplete gamma function ρ(x) = +(2πℓ2)−1Γ(N, |z|2)/Γ(N). +Constancy of density in the +bulk is then manifest, as is its drop to zero close to +the edge |z| = +√ +N, with an error function behavior +that can be deduced from known asymptotic formulas +for gamma functions [76] and reproduces Eqs. (14)–(54) +with λ = 2f ′ = 2. +Current. For the LLL states (3) with definite angular +momentum, each one-body current (55) is purely angu- +lar, i.e. it reads jm = (. . .)dϕ. The sum (56) can then +be evaluated in closed form owing to an exact cancella- +tion between the contribution of the states m and m+ 1, +eventually leading to a current only due to the (N − 1)th +wave function. It is then trivial to show that the current +is localized as a Gaussian close to the edge, since this is +also true of the underlying single-particle wave function. +This reproduces Eqs. (15)–(57) with λ = 2f ′ = 2. +Correlations. +The computation of electronic correla- +tions close to the edge is similar to that of the density. +Indeed, since the many-body ground state wave function +is a Slater determinant, the two-point correlation func- +tion in the ground state can be expressed as in (59). The +exact wave functions (3) can then be used to write the +correlation (59) as an incomplete gamma function (this +time with a complex argument): +C(z, ¯z, w, ¯w) = +1 +2πℓ2 +Γ(N, ¯zw) +Γ(N) +e−(|z|2+|w|2)/2 ez ¯ +w. (A4) +It is then manifest that bulk correlations coincide with +the kernel (5) at leading order in the thermodynamic +limit. As for the edge behavior, it can be extracted e.g. +from a steepest descent argument [76] and reproduces +Eqs. (16)–(61) with λ = 2f ′ = 2. +Appendix B: Semiclassical expansion of P V P +In this appendix, we derive (22) starting from (21). +To this end, think of V (x, y) as some smooth function of + +16 +(x, y) whose arguments can be complexified, and change +the integration variables (x, y) of (21) to +s ≡ x − +ℓ +√ +2(z + ¯w), +t ≡ y + +iℓ +√ +2(z − ¯w). +(B1) +In terms of (s, t), the integrals in (21) are two line in- +tegrals in the complex plane, each along a path from +−∞ + ic to +∞ + ic, where c is some irrelevant real con- +stant (a different one for s and t). The advantage of the +change of variables (B1) is to make the exponential factor +in (21) purely Gaussian: +⟨z, ¯z|P V (x)P|w, ¯w⟩ = +1 +(2πℓ2)2 e− |z−w|2 +2 +e +z ¯ +w−¯zw +2 +× +� +ds dt V +� +s+ ℓ +√ +2(z + ¯w), t− iℓ +√ +2(z − ¯w) +� +e− s2+t2 +2ℓ2 . +(B2) +We then complexify V , thus replacing V (x, y) by V(z, ¯z), +where V(z, ¯w) is a function of two complex variables, +holomorphic in the first one and anti-holomorphic in the +second. +We can then deform independently both inte- +gration contours for s and t back to the real line. For +small ℓ, the Gaussian factor of (B2) localizes everything +to s = t = 0. We now use our assumption of slow varia- +tion of V (x) to Taylor-expand it as +V +� +s + +ℓ +√ +2(z + ¯w), t − +iℓ +√ +2(z − ¯w) +� +∼ +� +V + s2 +2 ∂2 +xV + t2 +2 ∂2 +yV +����� ℓ +√ +2 (z+ ¯ +w),− iℓ +√ +2 (z− ¯ +w) +�, +(B3) +where we only kept terms that give non-zero contribu- +tions to the O(ℓ2) approximation of the integral (B2). +Note that everything is evaluated at (x, y) = ( ℓ +√ +2(z + +¯w), − iℓ +√ +2(z − ¯w)); in complex coordinates, this is just +the point (z, ¯w), so it is simpler to write the potential +as V(z, ¯w). Plugging the expansion (B3) into (B2) then +yields the result (22). +Appendix C: The transport equation +The purpose of this appendix is to derive the real and +imaginary parts of the transport equation in (32) and +(38), respectively, by imposing the eigenvalue equation +(7) based on our WKB ansatz (30) in the case of edge- +deformed droplets. The derivation relies on expanding +the energy and the potential as in (9) and (22). +It is +divided in two parts. First, we use the eigenvalue equa- +tion to derive the constraint (31), and let z belong to an +equipotential so that the whole equation boils down to +a 1D integral identity. Second, we show that the inte- +gral has a sharp saddle point in the large m limit; this +allows us to rephrase the integral constraint as a first- +order transport equation for the unknown function n. +1. +Evaluation along an equipotential +Using the wave functions (18)–(19) and the expansion +(22) of the potential along with the projector property +P 2 = P, the eigenvalue problem (7) reads +0 = +� +R2 +d2w +2πℓ2 e− |z−w|2 +2 ++ z ¯ +w−¯zw +2 +�� +V+ ℓ2 +2 ∇2V +���� +(z, ¯ +w)−Em +� +× +� +dθ n(θ) eimθ δ2� +w − +� +F(m, θ), G(m, θ) +�� +(C1) +up to O(ℓ4) corrections [95]. In the case of edge-deformed +traps, V(z, ¯w) is the bicomplex potential given in (28) +and the delta function localizes the whole integral over +w to a level curve (27) with K = m. Integrating over w +and changing the integration variable from θ = f(ϕ) to +ϕ yields Eq. (31). +Note that the structure of Eqs. (C1) and (31) is 0 = +e−|z|2/2 F(z) for a holomorphic function F(z), so setting +F(z) = 0 on a closed curve implies F(z) = 0 everywhere. +Accordingly, we will solve (C1)–(31) along the equipo- +tential (26) by fixing K = m and parametrizing +z = +� +mf ′(α) eiα, +α ∈ [0, 2π). +(C2) +This ensures that all three terms in the exponent of (31) +are of the same order O(m). Then (31) with the choice +(C2) and ϕ ≡ α + ε becomes +0 = +� π +−π +dε f ′(α + ε) n(f(α + ε)) exp +� +imf(α + ε) − 1 +2mf ′(α + ε) + m +� +f ′(α)f ′(α + ε) e−iε� +× +� +V +�� +mf ′(α) eiα, +� +mf ′(α + ε) e−i(α+ε)� ++ ℓ2 +2 ∇2V − E0 +m − ℓ2E1 +m +� +. +(C3) +This rewriting will allow us to carry out the integral thanks to the saddle-point approximation, obtained by expanding +all terms in powers of ε and leading to a differential equation for n(θ). + +17 +2. +Saddle point and transport equation +The saddle-point expansion of the integral (C3) is cumbersome but straightforward. The strategy is to expand all +factors in the integrand up to a suitable power of ε, then perform the resulting integrals of the form +� +dε ε# e−Cε2, +where C is some f-dependent coefficient [see e.g. (C5)]. The powers of ε involved are typically small, as higher powers +are suppressed in the classical limit [large m and ℓ2m = O(1)]. The fact that the argument of n(θ) also involves a +factor ε eventually converts the integral into a transport equation of the form n′(θ) ∝ n(θ) [see (C16)]. +We start with (C3) and first expand the exponential, then the potential with its Laplacian, and finally the simplest +f ′(ϕ)n(f(ϕ)) prefactor. For convenience, we introduce the notation +A ≡ f ′′ +f ′ , +B ≡ f ′′′ +f ′ +(C4) +for combinations of derivatives of f that often appear below; from now on, expressions of the form f or f ′, etc., are +all implicitly evaluated at α unless specified otherwise (so f ≡ f(α), f ′ ≡ f ′(α), etc.). Note for future reference the +useful relation A′ = B − A2. +The exponential. Using the notation (C4), one has +exp +� +imf(α + ε) − 1 +2mf ′(α + ε) + m +� +f ′(α)f ′(α + ε) e−iε� +∼ eimf+ 1 +2 mf ′ exp +� +− 1 +2mf ′ � +1 + A2 +4 +� +ε2� � +1 + mf ′ε3 � +i +6 − A +4 − iB +12 + iA2 +8 − AB +8 + A3 +16 +�� +(C5) +where the factor exp +� +imf + mf ′/2 +� +is ultimately irrelevant for the eigenvalue equation (C3), so we will not include +it in what follows. The main point of (C5) is to exhibit the leading Gaussian behavior exp +� +−(mf ′/2)(1 + A2/4)ε2� +of the integrand, which will eventually allow us to convert (C3) into a differential equation for the unknown function +n(θ). In fact, the same exponential term appears in the approximately Gaussian wave function (43). +The potential. We now turn to the expansions of the potential and of its Laplacian. As a first step, our task is to +expand the potential +V +�� +mf ′ eiα, +� +mf ′(α + ε) e−i(α+ε)� += V0 +� +� +�ℓ2m +√f ′ � +f ′(α + ε) e−iε +f ′ +� +1 +2i log +� √f ′ e2iα+iε +√ +f ′(α+ε) +�� +� +� +� +∼ V0 +� +ℓ2m +� +1 − iε +� +1 + A2 +4 +� ++ ε2 � +− 1 +2 + B +8 − 3A2 +8 +− A3 +4i − A4 +16 + AB +4i + A2B +32 +� �� +∼ V0 +� +ℓ2m +� +− iℓ2m ε +� +1 + A2 +4 +� +V ′ +0 +� +ℓ2m +� +− 1 +2ℓ4m2ε2 � +1 + A2 +4 +�2 +V ′′ +0 +� +ℓ2m +� ++ ℓ2m ε2 � +− 1 +2 + B +8 − 3A2 +8 +− A3 +4i − A4 +16 + AB +4i + A2B +32 +� +V ′ +0 +� +ℓ2m +� +, +(C6) +where we used (28) and then the notation (C4). Aside +from the contribution of the Laplacian, these are all the +terms of the potential needed in the eigenvalue equation +(C3) along an equipotential. As expected, they all ul- +timately involve the potential and its derivatives at the +equipotential (26). For ε = 0, the whole expression boils +down to V0(ℓ2m) alone. +Let us now turn to the Laplacian term. +The +eigenvalue +equation +(C3) +requires +the +Laplacian +evaluated +at +the +complexified +point +(z, ¯w) += +�� +mf ′(α) eiα, +� +mf ′(α + ε) e−i(α+ε)� +. +In +practice, +the Laplacian term is multiplied by ℓ2 in (C3), so we +may safely set ε = 0 when computing it; this removes +the complexification and allows us to write the Laplacian +contribution in (C3) as +ℓ2 +2 ∇2V ∼ ℓ2 +f ′ +� +1 − B +4 + A2 +2 +� +V ′ +0(ℓ2m) ++ ℓ4m +f ′ +� +1 + A2 +4 +� +V ′′ +0 (ℓ2m), +(C7) +which follows from the general expression (29) evaluated +on the equipotential (26). +All together. Let us finally consider the very first factor +on the right-hand side of (C3), namely +f ′(α + ε) n(f(α + ε)) ∼ f ′n(f) + ε +� +f ′′n(f) + f ′2n′(f) +� +(C8) +where higher powers of ε are negligible at this order. To +see why they may be neglected, it is helpful to investigate + +18 +the general structure of the small ℓ expansion of (C3): +the exponential term in (C5) has the form +exp[imf(. . .)] ∼ const × e−mΛε2(1 + mLε3) +(C9) +with m ≫ 1 and Λ, L some O(1) coefficients. +Sim- +ilarly, the potential expansion (C6) together with the +Laplacian correction (C7) can schematically be written as +V0+ ℓ2 +2 ∇2V0 ∼ V0+ℓ2W0+Gε+Hε2, where V0 ≡ V0(ℓ2m) +while W0, G, H are again some O(1) coefficients. Finally, +the expansion (C8) of the prefactor roughly has the form +f ′n e(...) ∼ const × (f ′n + εIn′ + εJn), +(C10) +where I, J are O(1) coefficients. +Putting together the +schematic expressions (C9)–(C10) and using the fact that +constant (i.e. ε-independent) contributions are irrelevant, +the eigenvalue equation (C3) becomes +0 = +� +dε (f ′n + εIn′ + εJn)e−mΛε2(1 + mLε3) +× +� +V0 + ℓ2W0 + Gε + Hε2 − E0 +m − ℓ2E1 +m +� +. +(C11) +Here the right-hand side is a sum of integrals whose in- +tegrand has the form εn e−mΛε2. For odd n, each such +integral vanishes; for even n, it is non-zero and scales as +m−n/2. This is why only the first order in ε is needed in +the expansion (C8): higher powers of ε would yield sub- +leading corrections to (C11), which can be consistently +taken into account only by expanding the exponential, +potential and Laplacian terms up to orders in ε higher +than what we did above. Here we content ourselves with +the zeroth and first order terms in ℓ2 (i.e. in 1/m). At +that level of approximation, (C11) yields the zeroth order +statement +V0 − E0 +m = 0 +(C12) +and the first-order result +f ′n +� +Λℓ2m[W0 −E1 +m]+ H +2 + 3LG +4Λ +� ++ G +2 +� +In′ +Jn +� += 0, +(C13) +where ℓ2m = O(1) as before. Eq. (C12) confirms that the +eigenvalue equation holds if E0 +m = V0(ℓ2m), i.e. if the en- +ergy of the eigenstate |ψm⟩ is that of its equipotential at +leading order [recall (9)]. More important, (C13) yields +a transport equation for n, whose schematic form is +GI +2 +n′ +n +f ′� +Λℓ2m(W0−E1 +m)+ H +2 + 3LG +4Λ +� ++ GJ +2 = 0. (C14) +We now rely on the expansions (C5)–(C8) to write this +transport equation explicitly: using the notation (33) and +plugging (C5)–(C8) into (C3) yields the condition +0 = +� +dε e +− Kf ′ +2 +� +1+ A2 +4 +� +ε2 � +1 + ε +� +A + f ′ n′(f) +n(f) +�� � +1 + Kf ′ε3 � +i +6 − A +4 − iB +12 + iA2 +8 − AB +8 + A3 +16 +�� +× +� +−iℓ2Kε +� +1 + A2 +4 +� +Ωm − ℓ2Kε2 +2 +� +1 + A2 +4 +�2 +Γm + ℓ2Kε2 � +− 1 +2 + B +8 − 3A2 +8 +− A3 +4i − A4 +16 + AB +4i + A2B +32 +� +Ωm ++ ℓ2 +f ′ +� +1 − B +4 + A2 +2 +� +Ωm + ℓ2 +f ′ +� +1 + A2 +4 +� +Γm − ℓ2E1 +m +� +, +(C15) +whose structure is that announced in (C11), as had to +be the case. What remains is to multiply all the factors +in the integrand, keep track of powers of ε and integrate +over ε, which leads to +iR′/R = +� +1 + A2 +4 +� +Γm +2Ωm + 1 − B +4 + A2 +2 − f ′ E1 +m +Ωm +− +1 +1 + A2 +4 +�� +B +8 + A4 +16 − A2B +32 +� ++ i +� +A +4 + 3A3 +16 − AB +8 +�� +, +(C16) +where we introduced R ≡ R(α) ≡ n(f(α)) for simplicity. +This is the transport equation for the O(1) multiplica- +tive factor of the WKB ansatz (30). Its real and imag- +inary parts, respectively, govern the phase and norm of +n(f(ϕ)) ≡ N(ϕ) eiΦ(ϕ): +−Φ′ = +� +1 + A2 +4 +� +Γm +2Ωm − f ′ E1 +m +Ωm ++ +1 +1+ A2 +4 +� +1 + 3A2 +4 ++ A4 +16 − 3B +8 − A2B +32 +� +, (C17) +N ′/N = − +1 +1 + A2 +4 +� +A +4 + 3A3 +16 − AB +8 +� +. +(C18) +The identity B = A′+A2 then reduces these two relations +to Eqs. (32) and (38) in the main text. + +19 +[1] K. von Klitzing, G. Dorda, and M. Pepper, New method +for high-accuracy determination of the fine-structure con- +stant based on quantized Hall resistance, Phys. Rev. Lett. +45, 494 (1980). +[2] D. C. Tsui, H. L. Stormer, and A. C. Gossard, Two- +dimensional magnetotransport in the extreme quantum +limit, Phys, Rev. Lett. 48, 1559 (1982). +[3] R. B. Laughlin, Anomalous quantum Hall effect: An in- +compressible quantum fluid with fractionally charged ex- +citations, Phys. Rev. Lett. 50, 1395 (1983). +[4] P. Delplace and G. Montambaux, WKB analysis of edge +states in graphene in a strong magnetic field, Phys. Rev. +B 82, 205412 (2010), arXiv:1007.2910 [cond-mat.mes- +hall]. +[5] G. Montambaux, Semiclassical quantization of skipping +orbits, Eur. Phys. J. B 79, 215 (2011), arXiv:1006.5819 +[cond-mat.mes-hall]. +[6] S. Bahcall and L. Susskind, Fluid dynamics, Chern- +Simons theory and the quantum Hall effect, Int. J. Mod. +Phys. B 5, 2735 (1991). +[7] J. Fröhlich and A. Zee, Large scale physics of the quan- +tum Hall fluid, Nucl. Phys. B 364, 517 (1991). +[8] X. G. Wen, Electrodynamical properties of gapless edge +excitations in the fractional quantum Hall states, Phys. +Rev. Lett. 64, 2206 (1990). +[9] X. G. Wen, Chiral Luttinger liquid and the edge excita- +tions in the fractional quantum Hall states, Phys. Rev. +B 41, 12838 (1990). +[10] X.-G. Wen, Theory of the edge states in fractional quan- +tum Hall effects, Int. J. Mod. Phys. B 06, 1711 (1992). +[11] P. Weitz, E. Ahlswede, J. Weis, K. Klitzing, and K. Eberl, +Hall-potential investigations under quantum Hall condi- +tions using scanning force microscopy, Physica E 6, 247 +(2000). +[12] E. Ahlswede, P. Weitz, J. Weis, K. von Klitzing, and +K. Eberl, Hall potential profiles in the quantum Hall +regime measured by a scanning force microscope, Physica +B 298, 562 (2001). +[13] J. Weis, Hall potential profiles in quantum Hall samples +measured by a low-temperature scanning force micro- +scope, Int. J. Mod. Phys. B 21, 1297 (2007). +[14] S. Ilani, J. Martin, E. Teitelbaum, J. H. Smet, D. Mahalu, +V. Umansky, and A. Yacoby, The microscopic nature of +localization in the quantum Hall effect, Nature 427, 328 +(2004), arXiv:cond-mat/0402026. +[15] G. A. Steele, R. C. Ashoori, L. N. Pfeiffer, and K. W. +West, Imaging transport resonances in the quantum Hall +effect, Phys. Rev. Lett. 95, 136804 (2005), arXiv:cond- +mat/0506347. +[16] A. Kamiyama, M. Matsuura, J. N. Moore, T. Mano, +N. Shibata, and G. Yusa, Real-time and space visual- +ization of excitations of the ν = 1/3 fractional quan- +tum Hall edge, Phys. Rev. Research 4, L012040 (2022), +arXiv:2201.10180 [cond-mat.mes-hall]. +[17] V. Bretin, S. Stock, Y. Seurin, and J. Dalibard, Fast +rotation of a Bose-Einstein condensate, Phys. Rev. Lett. +92, 050403 (2004). +[18] V. Schweikhard, I. Coddington, P. Engels, V. P. Mogen- +dorff, and E. A. Cornell, Rapidly rotating Bose-Einstein +condensates in and near the lowest Landau level, Phys. +Rev. Lett. 92, 040404 (2004), arXiv:cond-mat/0308582. +[19] N. R. Cooper, Rapidly rotating atomic gases, Adv. Phys. +57, 539 (2008), arXiv:0810.4398 [cond-mat.mes-hall]. +[20] B. Dellabetta, T. L. Hughes, M. J. Gilbert, and B. L. Lev, +Imaging topologically protected transport with quan- +tum degenerate gases, Phys. Rev. B 85, 205442 (2012), +arXiv:1202.0060 [cond-mat.mtrl-sci]. +[21] N. Goldman, J. Beugnon, and F. Gerbier, Identifying +topological edge states in 2D optical lattices using light +scattering, Eur. Phys. J. Spec. Top. 217, 135 (2013), +arXiv:1209.2051 [cond-mat.quant-gas]. +[22] N. Goldman, J. Dalibard, A. Dauphin, F. Gerbier, +M. Lewenstein, P. Zoller, and I. B. Spielman, Direct +imaging of topological edge states in cold-atom systems, +Proc. Natl. Acad. Sci. 110, 6736 (2013), arXiv:1212.5093 +[cond-mat.quant-gas]. +[23] R. J. Fletcher, A. Shaffer, C. C. Wilson, P. B. Pa- +tel, Z. Yan, V. Crépel, B. Mukherjee, and M. W. +Zwierlein, Geometric squeezing into the lowest Landau +level, Science 372, 1318 (2021), arXiv:1911.12347 [cond- +mat.quant-gas]. +[24] B. Mukherjee, A. Shaffer, P. B. Patel, Z. Yan, C. C. Wil- +son, V. Crépel, R. J. Fletcher, and M. Zwierlein, Crys- +tallization of bosonic quantum Hall states in a rotating +quantum gas, Nature 601, 58 (2022), arXiv:2106.11300 +[cond-mat.quant-gas]. +[25] B. I. Halperin, Quantized Hall conductance, current- +carrying edge states, and the existence of extended states +in a two-dimensional disordered potential, Phys. Rev. B +25, 2185 (1982). +[26] S. A. Trugman, Localization, percolation, and the quan- +tum Hall effect, Phys. Rev. B 27, 7539 (1983). +[27] R. Joynt and R. E. Prange, Conditions for the quantum +Hall effect, Phys. Rev. B 29, 3303 (1984). +[28] T. Champel and S. Florens, Quantum transport proper- +ties of two-dimensional electron gases under high mag- +netic fields, Phys. Rev. B 75, 245326 (2007), arXiv:cond- +mat/0612149. +[29] T. Champel, S. Florens, and L. Canet, Microscopics +of disordered two-dimensional electron gases under high +magnetic fields: Equilibrium properties and dissipation +in the hydrodynamic regime, Phys. Rev. B 78, 125302 +(2008), arXiv:0806.3038 [cond-mat.mes-hall]. +[30] P. Choquard, B. Piller, and R. Rentsch, On the dielectric +susceptibility of classical Coulomb systems. II, J. Stat. +Phys. 46, 599 (1987). +[31] B. Jancovici, Classical Coulomb systems: Screening and +correlations revisited, J. Stat. Phys. 80, 445 (1995), +arXiv:cond-mat/9503109. +[32] A. Zabrodin and P. Wiegmann, Large-N expansion for +the 2D Dyson gas, J. Phys. A: Math. Gen. 39, 8933 +(2006), arXiv:hep-th/0601009. +[33] B. Bradlyn, M. Goldstein, and N. Read, Kubo formulas +for viscosity: Hall viscosity, Ward identities, and the re- +lation with conductivity, Phys. Rev. B 86, 245309 (2012), +arXiv:1207.7021 [cond-mat.stat-mech]. +[34] B. Bradlyn and N. Read, Topological central charge from +Berry curvature: Gravitational anomalies in trial wave +functions for topological phases, Phys. Rev. B 91, 165306 +(2015), arXiv:1502.04126 [cond-mat.mes-hall]. +[35] A. G. Abanov and A. Gromov, Electromagnetic and +gravitational responses of two-dimensional noninteract- + +20 +ing electrons in a background magnetic field, Phys. Rev. +B 90, 014435 (2014), arXiv:1401.3703 [cond-mat.str-el]. +[36] T. Can, M. Laskin, and P. Wiegmann, Fractional quan- +tum Hall effect in a curved space: gravitational anomaly +and electromagnetic response, Phys. Rev. Lett. 113, +046803 (2014), arXiv:1402.1531 [cond-mat.str-el]. +[37] T. Can, M. Laskin, and P. B. Wiegmann, Geome- +try of quantum Hall states: +Gravitational anomaly +and transport coefficients, Ann. Phys. 362, 752 (2015), +arXiv:1411.3105 [cond-mat.str-el]. +[38] T. Can, Y. Chiu, M. Laskin, and P. Wiegmann, Emer- +gent conformal symmetry and geometric transport prop- +erties of quantum Hall states on singular surfaces, Phys. +Rev. Lett. 117, 266803 (2016), arXiv:1602.04802 [cond- +mat.str-el]. +[39] S. Klevtsov and P. Wiegmann, Geometric adiabatic +transport in quantum Hall states, Phys. Rev. Lett. 115, +086801 (2015), arXiv:1504.07198 [cond-mat.str-el]. +[40] S. M. Girvin and T. Jach, Formalism for the quantum +Hall effect: +Hilbert space of analytic functions, Phys. +Rev. B 29, 5617 (1984). +[41] M. R. Geller and G. Vignale, Currents in the compress- +ible and incompressible regions of the two-dimensional +electron gas, Phys. Rev. B 50, 11714 (1994). +[42] M. R. Geller and G. Vignale, Universal equilibrium cur- +rents in the quantum Hall fluid, Phys. Rev. B 52, 14137 +(1995), arXiv:cond-mat/9412028. +[43] As usual, the semiclassical limit is singular and involves +dimensionful quantities. The limit ℓ → 0 thus really +means that all length scales of interest are much larger +than the magnetic length. +[44] J. Bellissard, A. van Elst, and H. Schulz-Baldes, The +noncommutative geometry of the quantum Hall effect, J. +Math. Phys. 35, 5373 (1994), arXiv:cond-mat/9411052. +[45] S. Girvin, A. MacDonald, and P. Platzman, Magneto- +roton theory of collective excitations in the fractional +quantum Hall effect, Phys. Rev. B 33, 2481 (1986). +[46] V. Pasquier, Quantum Hall effect and non-commutative +geometry, in Quantum Spaces, Progress in Mathemati- +cal Physics, Vol. 53, edited by B. Duplantier and V. Ri- +vasseau (Birkhäuser, Basel, 2007) pp. 1–17. +[47] S. Iso, D. Karabali, and B. Sakita, Fermions in the low- +est Landau level: Bosonization, W∞ algebra, droplets, +chiral bosons, Phys. Lett. B 296, 143 (1992), arXiv:hep- +th/9209003. +[48] L. Susskind, The quantum Hall fluid and noncommuta- +tive Chern-Simons theory (2001), arXiv:hep-th/0101029. +[49] A. P. Polychronakos, Quantum Hall states as matrix +Chern-Simons theory, JHEP 2001 (04), 011, arXiv:hep- +th/0103013. +[50] S. Hellerman and M. Van Raamsdonk, Quantum Hall +physics = noncommutative field theory, JHEP 2001 (10), +039, arXiv:hep-th/0103179. +[51] E. Fradkin, V. Jejjala, and R. G. Leigh, Noncommutative +Chern-Simons for the quantum Hall system and duality, +Nucl. Phys. B 642, 483 (2002), arXiv:cond-mat/0205653. +[52] Z. Dong and T. Senthil, Noncommutative field theory +and composite Fermi liquids in some quantum Hall sys- +tems, Phys. Rev. B 102, 205126 (2020), arXiv:2006.01282 +[cond-mat.str-el]. +[53] Y.-H. Du, U. Mehta, and D. T. Son, Noncommutative +gauge symmetry in the fractional quantum Hall effect +(2021), arXiv:2110.13875 [cond-mat.str-el]. +[54] L. Charles, Berezin-Toeplitz operators, a semi-classical +approach, Commun. Math. Phys. 239, 1 (2003). +[55] L. Charles, Quasimodes and Bohr-Sommerfeld conditions +for the Toeplitz operators, Commun. Partial. Differ. Equ. +28, 1527 (2003). +[56] C. M. Bender and S. A. Orszag, Advanced Mathemati- +cal Methods for Scientists and Engineers I: Asymptotic +Methods and Perturbation Theory (Springer, New York, +1999). +[57] B. Oblak and B. Estienne, Adiabatic deformations of +quantum Hall droplets (2022), arXiv:2212.12935 [cond- +mat.mes-hall]. +[58] A. Cappelli, C. A. Trugenberger, and G. R. Zemba, In- +finite symmetry in the quantum Hall effect, Nucl. Phys. +B 396, 465 (1993), arXiv:hep-th/9206027. +[59] A. Cappelli, G. V. Dunne, C. A. Trugenberger, and G. R. +Zemba, Conformal symmetry and universal properties of +quantum Hall states, Nucl. Phys. B 398, 531 (1993), +arXiv:hep-th/9211071. +[60] A. Cappelli, C. A. Trugenberger, and G. R. Zemba, Clas- +sification of quantum Hall universality classes by W1+∞ +symmetry, Phys. Rev. Lett. 72, 1902 (1994), arXiv:hep- +th/9310181. +[61] A. Cappelli, C. A. Trugenberger, and G. R. Zemba, +W1+∞ dynamics of edge excitations in the quantum +Hall effect, Ann. Phys. 246, 86 (1996), arXiv:cond- +mat/9407095. +[62] A. Cappelli, C. A. Trugenberger, and G. R. Zemba, +Quantum Hall fluids as W1+∞ minimal models, in Par- +ticles and Fields, CRM Series in Mathematical Physics, +edited by G. Semenoff and L. Vinet (Springer, New York, +1999) arXiv:hep-th/9502050. +[63] A. Cappelli, C. A. Trugenberger, and G. R. Zemba, +W1+∞ field theories for the edge excitations in the quan- +tum Hall effect, Int. J. Mod. Phys. A 12, 1101 (1997), +arXiv:hep-th/9610019. +[64] A. Izosimov, B. Khesin, and M. Mousavi, Coadjoint +orbits of symplectic diffeomorphisms of surfaces and +ideal hydrodynamics, Ann. Inst. Fourier 66, 2385 (2016), +arXiv:1504.05629 [math.SG]. +[65] A. C. da Silva, Lectures on Symplectic Geometry, Lec- +ture Notes in Mathematics, Vol. 1764 (Springer, Berlin, +Heidelberg, 2008). +[66] It is more common to write the action variable as I rather +than K, but the notation here is more convenient for later +computations involving exponentials of i times an action. +[67] B. Jancovici, Exact results for the two-dimensional one- +component plasma, Phys. Rev. Lett. 46, 386 (1981). +[68] P. J. Forrester and B. Jancovici, Two-dimensional one- +component plasma in a quadrupolar field, Int. J. Mod. +Phys. A 11, 941 (1996), arXiv:cond-mat/9509111. +[69] B. Rider, A limit theorem at the edge of a non-Hermitian +random matrix ensemble, J. Phys. A: Math. Gen. 36, +3401 (2003). +[70] M. Bender, Edge scaling limits for a family of non- +Hermitian random matrix ensembles, Probab. The- +ory Relat. Fields 147, 241 (2010), arXiv:0808.2608 +[math.PR]. +[71] S.-Y. Lee and R. Riser, Fine asymptotic behavior for +eigenvalues of random normal matrices: Ellipse case, J. +Math. Phys. 57, 023302 (2016), arXiv:1501.02781 [math- +ph]. +[72] This is ‘harmonic’ in that it holds when V0(r2/2) = +Ω r2/2 in Eq. (25). + +21 +[73] One finds δ2 = ia2 � +2−{f(α),α} +4 +− i f′′(α) +2f′(α) +� +1 − i f′′(α) +2f′(α) +�� +× +� +1−i f′′(α) +2f′(α) +�−3 +with the Schwarzian derivative {f(α), α}. +[74] P. Di Francesco, M. Gaudin, C. Itzykson, and F. Lesage, +Laughlin’s wave functions, Coulomb gases and expan- +sions of the discriminant, Int. J. Mod. Phys. A 09, 4257 +(1994), arXiv:hep-th/9401163. +[75] Properly +normalized, +the +tangent +vector +is +�√ +2ℓ/σ√f ′�� +r−1∂ϕ + (f ′′/2f ′)∂r +� +for |r| = ℓ√2Nf ′. We +recall that dϕ · ∂ϕ = 1 = dr · ∂r and dϕ · ∂r = 0 = dϕ · ∂r +with ||∂ϕ|| = r = ||dϕ||−1 and ||∂r|| = 1 = ||dr||. +[76] B. Estienne, B. Oblak, and J.-M. Stéphan, Ergodic edge +modes in the 4D quantum Hall effect, SciPost Phys. 11, +016 (2021), arXiv:2104.01860 [cond-mat.mes-hall]. +[77] P. Wiegmann and A. Zabrodin, Large scale correlations +in normal non-Hermitian matrix ensembles, J. Phys. A: +Math. Gen. 36, 3411 (2003), arXiv:hep-th/0210159. +[78] P. Wiegmann, Nonlinear hydrodynamics and fractionally +quantized solitons at the fractional quantum Hall edge, +Phys. Rev. Lett. 108, 206810 (2012), arXiv:1112.0810 +[cond-mat.mes-hall]. +[79] A. Nardin and I. Carusotto, Non-linear edge dynamics +of an integer quantum Hall fluid, Europhys. Lett. 132, +10002 (2020), arXiv:2007.12945 [cond-mat.quant-gas]. +[80] A. Nardin and I. Carusotto, Linear and nonlinear edge +dynamics of trapped fractional quantum Hall droplets +beyond the chiral Luttinger liquid paradigm (2022), +arXiv:2203.02539 [cond-mat.mes-hall]. +[81] F. D. M. Haldane, ‘Luttinger liquid theory’ of one- +dimensional quantum fluids. I. Properties of the Lut- +tinger model and their extension to the general 1D inter- +acting spinless Fermi gas, J. Phys. C: Solid State Phys. +14, 2585 (1981). +[82] N. Allegra, J. Dubail, J.-M. Stéphan, and J. Viti, Inho- +mogeneous field theory inside the arctic circle, J. Stat. +Mech. 2016, 053108 (2016), arXiv:1512.02872 [cond- +mat.stat-mech]. +[83] J. Dubail, J.-M. Stéphan, J. Viti, and P. Calabrese, Con- +formal field theory for inhomogeneous one-dimensional +quantum systems: The example of non-interacting Fermi +gases, SciPost Phys. 2, 002 (2017), arXiv:1606.04401 +[cond-mat.str-el]. +[84] J. Dubail, J.-M. Stéphan, and P. Calabrese, Emergence of +curved light-cones in a class of inhomogeneous Luttinger +liquids, SciPost Phys. 3, 019 (2017), arXiv:1705.00679 +[cond-mat.str-el]. +[85] K. Gawędzki, E. Langmann, and P. Moosavi, Finite-time +universality in nonequilibrium CFT, J. Stat. Phys. 172, +353 (2018), arXiv:1712.00141 [cond-mat.stat-mech]. +[86] E. Langmann and P. Moosavi, Diffusive heat waves in +random conformal field theory, Phys. Rev. Lett. 122, +020201 (2019), arXiv:1807.10239 [cond-mat.stat-mech]. +[87] P. +Moosavi, +Inhomogeneous +conformal +field +theory +out +of +equilibrium, +Ann. +Henri +Poincaré +(2021), +arXiv:1912.04821 [math-ph]. +[88] M. Gluza, P. Moosavi, and S. Sotiriadis, Breaking of +Huygens-Fresnel principle in inhomogeneous Tomonaga- +Luttinger liquids, J. Phys. A: Math. Theor. 55, 054002 +(2022), arXiv:2104.07751 [cond-mat.stat-mech]. +[89] P. Moosavi, Exact Dirac-Bogoliubov-de Gennes dy- +namics +for +inhomogeneous +quantum +liquids +(2022), +arXiv:2208.14467 [cond-mat.stat-mech]. +[90] V. I. Arnold, Sur la géométrie différentielle des groupes +de +Lie +de +dimension +infinie +et +ses +applications +à +l’hydrodynamique des fluides parfaits, Ann. Inst. Fourier +16, 319 (1966). +[91] V. I. Arnold and B. A. Khesin, Topological Methods +in Hydrodynamics, Applied Mathematical Sciences, Vol. +125 (Springer, New York, 1998). +[92] B. Oblak and G. Kozyreff, Berry phases in the re- +constructed KdV equation, Chaos 30, 113114 (2020), +arXiv:2002.01780 [math-ph]. +[93] L. V. Delacrétaz, Y.-H. Du, U. Mehta, and D. T. Son, +Nonlinear bosonization of Fermi surfaces: The method of +coadjoint orbits, Phys. Rev. Research 4, 033131 (2022), +arXiv:2203.05004 [cond-mat.str-el]. +[94] G. W. Moore and N. Read, Nonabelions in the fractional +quantum Hall effect, Nucl. Phys. B 360, 362 (1991). +[95] Strictly speaking, the left-hand side of (C1) should be +O(ℓ4) since we already truncated the expansion of the +P V P operator to order O(ℓ2), but we just write it as +zero for simplicity. +