diff --git "a/0dFQT4oBgHgl3EQfDjUj/content/tmp_files/2301.13234v1.pdf.txt" "b/0dFQT4oBgHgl3EQfDjUj/content/tmp_files/2301.13234v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/0dFQT4oBgHgl3EQfDjUj/content/tmp_files/2301.13234v1.pdf.txt" @@ -0,0 +1,10427 @@ +arXiv:2301.13234v1 [cond-mat.str-el] 30 Jan 2023 +Field Theoretic Aspects of Condensed Matter Physics: An +Overview⋆ +Eduardo Fradkin +Department of Physics and Institute for Condensed Matter Theory, +University of Illinois at Urbana-Champaign, 1110 West Green St, Urbana Illinois 61801-3080, U.S.A. +Abstract +In this chapter I discuss the impact of concepts of Quantum Field Theory in modern Condensed +Physics. Although the interplay between these two areas is certainly not new, the impact and +mutual cross-fertilization has certainly grown enormously with time, and quantum Field Theory +has become a central conceptual tool in Condensed Matter Physics. In this chapter I cover how +these ideas and tools have influenced our understanding of phase transitions, both classical and +quantum, as well as topological phases of matter, and dualities. +Keywords: +Contents +1 +Introduction +3 +2 +Early Years: Feynman Diagrams and Correlation Functions +3 +3 +Critical Phenomena +4 +3.1 +Classical Critical Phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . . +4 +3.2 +Landau-Ginzburg Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +4 +3.3 +The Renormalization Group +. . . . . . . . . . . . . . . . . . . . . . . . . . . . +5 +3.3.1 +Scaling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +5 +3.3.2 +The Operator Product Expansion . . . . . . . . . . . . . . . . . . . . . . +6 +3.3.3 +Fixed Points +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +6 +3.3.4 +Universality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +7 +3.3.5 +RG flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +8 +3.3.6 +Asymptotic Freedom . . . . . . . . . . . . . . . . . . . . . . . . . . . . +9 +4 +Quantum Criticality +11 +4.1 +Dynamic Scaling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +12 +4.2 +The Ising Model in a Transverse Field . . . . . . . . . . . . . . . . . . . . . . . +12 +4.3 +Quantum Antiferromagnets and Non-Linear Sigma Models . . . . . . . . . . . . +14 +4.3.1 +Spin coherent states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +14 +4.3.2 +Path integral for a spin-S degree of freedom . . . . . . . . . . . . . . . . +15 +4.3.3 +Quantum Ferromagnet . . . . . . . . . . . . . . . . . . . . . . . . . . . +16 +4.3.4 +Quantum Antiferromagnet . . . . . . . . . . . . . . . . . . . . . . . . . +16 +5 +Topological Excitations +17 +5.1 +Topological Excitations: Vortices and Magnetic Monopoles . . . . . . . . . . . . +18 +5.1.1 +Vortices in two dimensions . . . . . . . . . . . . . . . . . . . . . . . . . +18 +5.1.2 +Magnetic monopoles in compact electrodynamics . . . . . . . . . . . . . +22 +5.2 +Non-Linear Sigma Models and Antiferromagnetic Quantum Spin Chains . . . . . +25 +5.3 +Topology and open integer-spin chains . . . . . . . . . . . . . . . . . . . . . . . +26 +⋆Work was supported in part by the US National Science Foundation through grant No. DMR 1725401 at the Uni- +versity of Illinois. +Preprint submitted to Encyclopedia of Condensed Matter Physics 2e +February 1, 2023 + +6 +Duality in Ising Models +27 +6.1 +Duality in the 2D Ising Model +. . . . . . . . . . . . . . . . . . . . . . . . . . . +27 +6.2 +The 3D duality: Z2 gauge theory . . . . . . . . . . . . . . . . . . . . . . . . . . +28 +7 +Bosonization +31 +7.1 +Dirac fermions in one space dimensions . . . . . . . . . . . . . . . . . . . . . . +31 +7.2 +Chiral symmetry and chiral symmetry breaking . . . . . . . . . . . . . . . . . . +34 +7.3 +The chiral anomaly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +36 +7.4 +Bosonization, anomalies and duality . . . . . . . . . . . . . . . . . . . . . . . . +38 +8 +Fractional Charge +42 +8.1 +Solitons in one dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +42 +8.2 +Polyacetylene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +43 +8.3 +Fractionally charged solitons . . . . . . . . . . . . . . . . . . . . . . . . . . . . +43 +9 +Fractional Statistics +45 +9.1 +Basics of fractional statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . +46 +9.2 +What is a topological field theory . . . . . . . . . . . . . . . . . . . . . . . . . . +46 +9.3 +Chern-Simons Gauge Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . +47 +9.4 +BF gauge theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +48 +9.5 +Quantization of Abelian Chern-Simons Gauge Theory . . . . . . . . . . . . . . . +48 +9.6 +Vacuum degeneracy a torus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +50 +9.7 +Fractional Statistics and Braids . . . . . . . . . . . . . . . . . . . . . . . . . . . +51 +10 Topological Phases of Matter +54 +10.1 Topological Insulators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +54 +10.1.1 Dirac fermions in 2+1 dimensions . . . . . . . . . . . . . . . . . . . . . +54 +10.1.2 Dirac Fermions and Topological Insulators +. . . . . . . . . . . . . . . . +55 +10.1.3 Chern invariants +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +56 +10.1.4 The quantum Hall effect on a lattice . . . . . . . . . . . . . . . . . . . . +58 +10.1.5 The Anomalous Quantum Hall Effect . . . . . . . . . . . . . . . . . . . +61 +10.1.6 The Parity Anomaly +. . . . . . . . . . . . . . . . . . . . . . . . . . . . +62 +10.2 Three-dimensional Z2 topological insulators . . . . . . . . . . . . . . . . . . . . +67 +10.2.1 Z2 Topological Invariants +. . . . . . . . . . . . . . . . . . . . . . . . . +67 +10.2.2 The Axial Anomaly and the Effective Action +. . . . . . . . . . . . . . . +68 +10.2.3 Theta terms, and Domain walls: Anomaly and the Callan-Harvey Effect . +71 +10.3 Chern-Simons Gauge Theory and The Fractional Quantum Hall Effect . . . . . . +73 +10.3.1 Landau levels and the Integer Hall effect . . . . . . . . . . . . . . . . . . +73 +10.3.2 The Laughlin Wave Function . . . . . . . . . . . . . . . . . . . . . . . . +77 +10.3.3 Quasiholes have fractional charge . . . . . . . . . . . . . . . . . . . . . +78 +10.3.4 The Jain States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +79 +10.3.5 Quasiholes have fractional statistics . . . . . . . . . . . . . . . . . . . . +80 +10.3.6 Hydrodynamic Effective Field Theory . . . . . . . . . . . . . . . . . . . +80 +10.3.7 Composite Boson Field Theory +. . . . . . . . . . . . . . . . . . . . . . +82 +10.3.8 Composite Fermion Field Theory +. . . . . . . . . . . . . . . . . . . . . +85 +10.3.9 The Compressible States . . . . . . . . . . . . . . . . . . . . . . . . . . +90 +10.3.10 Fractional Quantum Hall Wave Functions and Conformal Field Theory +. +91 +10.3.11 Edge States and Chiral Conformal Field Theory . . . . . . . . . . . . . . +99 +11 Particle-Vortex Dualities in 2+1 dimensions +105 +11.1 Electromagnetic Duality +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 +11.2 Particle-Vortex Duality in 2+1 dimensions . . . . . . . . . . . . . . . . . . . . . 105 +11.2.1 The 3D XY Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 +11.2.2 Scalar QED in 3D +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 +11.2.3 The duality mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 +11.3 Bosonization in 2+1 dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . 109 +11.3.1 Bosonization of the Dirac theory in 2+1 dimensions +. . . . . . . . . . . 110 +11.3.2 Bosonization of the Fermi Surface . . . . . . . . . . . . . . . . . . . . . 114 +12 Conclusions +116 +2 + +1. Introduction +Many (if not most) puzzling problems in Condensed Matter Physics involve systems with +a macroscopically large number of degrees of freedom often in regimes of large fluctuations, +thermal and/or quantum mechanical. The description of the physics of systems of this type +requires the framework provided by Quantum Field Theory. Although quantum field theory has +its origins in high-energy physics, notably in the development of Quantum Electrodynamics, it +has found a nurturing home in Condensed Matter Physics. +There is a long history of of cross-fertilization between both fields. Since the 1950’s in many +of the most significant developments in Condensed Matter Physics, Quantum Field Theory has +played a key role if not in the original development but certainly in the eventual understanding the +meaning and the further development of the discoveries. As a result, many of the discoveries and +concepts developed in Condensed Matter have had a reciprocal impact in Quantum Field theory. +One can already see this interplay in the development of the Bardeen-Cooper-Schrieffer theory +of superconductivity [1] and its implications in the theory of dynamical symmetry breaking in +particle physics by Nambu [2]. +The close and vibrant relationship between both fields has continued to these days, and it +is even stronger today than before. Many textbooks have been devoted to teaching these ideas +and concepts to new generations of condensed matter physicists (and field theorists as well). +The earlier texts focused on Green functions which are computed in perturbation theory using +Feynman diagrams [3] [4] [5], while the more modern ones have a broader scope, use path +integrals and attack non-perturbative problems [6, 7] [8] [9, 10] [11]. In two recent books I have +discussed many aspects of the interrelation between condensed matter physics and quantum field +theory in more depth than I can do in this chapter [9, 10]. +2. Early Years: Feynman Diagrams and Correlation Functions +Quantum field theory played a key role in the development of the Theory of the Fermi Liquid +[12, 13]. The theory of the Fermi liquid was first formulated by Landau using the framework of +hydrodynamics and the quantum Boltzmann equation. Landau’s ideas were later given a micro- +scopic basis using Green functions and Feynman diagrams [3], including the effects of quantum +fluctuations at finite temperature and non-equilibrium behavior [14, 15]. Linear response theory +was developed which allowed the computation of response functions (such as electrical conduc- +tivities and magnetic susceptibilities) from the computation of correlation functions for a given +microscopic theory [16]. In turn, correlation functions can be computed in terms of a set of Feyn- +man diagrams. These concepts and tools borrowed many concepts from field theory including +the study of the analytic structure of the generalized susceptibilities and the associated spectral +functions (together with the use of dispersion relations). These developments led to the deriva- +tion of the fluctuation-dissipation theorem. These ideas were widely applied to metals [13] and +superconductors [1], as well as to quantum magnets [17]. +The spectrum of an interacting system has low energy excitations characterized by a set of +quantum numbers associated with the symmetries of the theory. These low energy excitations are +known as quasiparticles. In the case of the Landau theory of the Fermi liquid the quasiparticle is +a “dressed” electron: it is a low energy excitation with the same quantum numbers (charge and +spin) and an electron but with a renormalized effective mass. There are many such quasiparticles +in condensed matter physics. The correlation functions (the propagators) of a physical system +has a specific analytic structure. In momentum (and frequency) space, the quasiparticle spectrum +is given by the poles of the correlators. +The role of symmetries and, in particular of gauge invariance, in the structure of correlation +functions was investigated extensively. A direct consequence of symmetries is the existence of +Ward identities which must be satisfied by all the correlation functions of the theory. Ward iden- +tities are exact relations that relate different correlation functions. Such identities contain a host +of important results. For example, in a theory with a globally conserved charge, the Hamiltonian +(and the action) have a global U(1) symmetry associated with the transformation of the local field +operator φ(x) (which can be fermionic or bosonic) to a new field φ′(x) = eiθφ(x) (where θ is a +constant phase). Theories with a global continuous symmetry have a locally conserved current +(and satisfy a continuity equation). The Ward identity requires the correlators of these currents +(and densities) to be be transverse (i.e. they should have vanishing divergence). In the absence +of so-called quantum anomalies (which we will discuss below) global symmetries can be made +local and become gauge symmetries. +3 + +In many circumstances a global symmetry can be spontaneously broken. If the global sym- +metry is continuous, then the Ward identities imply the existence of gapless excitations known +as Goldstone bosons. For example in the case of a superfluid, which has a spontaneously broken +U(1) symmetry the Goldstone boson is the gapless phase mode. Instead, a the N´eel phase of a +quantum antiferromagnet has two gapless Goldstone bosons, the magnons of the spontaneously +broken SO(3) global symmetry of this state of matter. Another example is the Ward identity of +quantum electrodynamics, which relates the electron self-energy to the electron-photon vertex +function, which also holds in non-relativistic electron fluids. In addition, these identities implied +the existence of sum rules that the spectral functions must satisfy. All of these results became part +of the standard toolkit of condensed matter experimentalists in analyzing their data and for the- +orists to make predictions. Ward identities and sum rules also imply restrictions on the allowed +approximations which are often needed to obtain predictions from a microscopic model. +3. Critical Phenomena +3.1. Classical Critical Phenomena +The late 1960s and particularly 1970s brought about an intense back and forth between con- +densed matter physics and field theory in the context of the problem of classical critical phenom- +ena and phase transitions. This was going to become a profound revolution on the description of +macroscopic physical systems with large-scale fluctuations. The problem of continuous (“sec- +ond order”) phase transitions has a long history going back to the work of Landau [18, 19] who +introduced the concept of an order parameter field. This turned out to be a powerful concept of +broad applicability in many physical systems sometimes quite different from each other at the +microscopic level. +A simple example is that of a ferromagnet with uniaxial anisotropy in which the spins of the +atoms in a crystal are strongly favored to be aligned (or anti-aligned) along certain directions of +the crystal. The simplest microscopic model for this problem is the Ising model, a spin system in +which the individual spins are allowed to take only two values, σ = ±1. The partition function +of the Ising model (in any dimension) is +Z = +� +[σ] +exp +− J +T +� +⟨r,r′⟩ +σ(r)σ(r′) + +(1) +where J is the exchange coupling constant and T is the temperature (measured in energy units); +here [σ] denotes the sume over the 2N spin configurations (for a lattice with N sites), and ⟨r, r′⟩ +are nearest neighbor sites of the lattice. The order parameter of the Ising model is the local mag- +netization ⟨σ(r)⟨ which, in the case of a ferromagnet, is uniform. The partition function of the +Ising model can be computed trivially in one dimension. The solution of the two-dimensional +Ising model by Onsager constituted a tour-de-force in theoretical physics [20]. Its actual meaning +remained obscure for some time. The work of Schultz, Mattis and Lieb [21] evinced a deep con- +nection between Onsager’s solution and the problem if the spectrum of one-dimensional quantum +spin chains [22] (specifically the one-dimensional Ising model in a transverse field [23]). One +important result was that the Ising model was in fact a theory of (free) fermions which, crudely +speaking, represented the configurations of domain walls of the magnet. However, even this sim- +ple model cannot be solved exactly in general dimension, and approximate mean field theories +of various sorts were devised over time to understand its physics. +3.2. Landau-Ginzburg Theory +Landau’s approach assumed that close enough to a phase transition, the important spin con- +figurations are those for which the local magnetization varies slowly on lattice scales. In this +picture the local magnetization, on long enough length scales, becomes an order parameter field +that takes values on the real numbers, and can be positive of negative. Thus, the order parameter +field is effectively the average of local magnetizations on some scale large compared to the lattice +scale, which we will denote by a real field φ(x). The thermodynamics properties of a system of +this type in d dimensions can be described in terms of a free energy +F[φ] = +� +ddx +�κ +2 (▽φ(x))2 + a(T − Tc)φ2(x) + uφ4(x) + . . . +� +(2) +4 + +which is known as the Ginzburg-Landaufree energy. Here κ is the stiffness of the order parameter +field, Tc is the (mean-field) critical temperature; a and u are two (positive) constants. This +expression make sense if the transition is continuous and hence that the order parameter is small +near the transition. The energy of the Ising model is invariant under the global symmetry [σ] �→ +[−σ]. This is the symmetry of the group Z2. Likewise, the Ginzburg-Landau free energy has the +global (discrete) symmetry [φ(x)] �→ [−φ(x)], and also has a Z2 global symmetry. +In Landau’s approach, which was a mean field theory, the equilibrium state is the global +minimum of this free energy. The nature of the equilibrium state depends on whether T > Tc +or T < Tc: for T > Tc the global minimum is the trivial configuration, ¯φ(x) = 0 (this is the +paramagnetic state), whereas for T < Tc the equilibrium state is two fold degenerate, ¯φ(x) = +±(a(Tc − T)/2)β, with the two degenerate states being related by the Z2 symmetry (this is the +ferromagnetic state). In the Landau theory the critical exponent of the magnetization is β = 1/2 +and the critical exponent of the correlation length is ν = 1/2. However, in the case of the 2D Ising +model the order parameter exponent is β = 1/8 [24] and the correlation length exponent is ν = 1. +These (and other) apparent discrepancies led many theorists for much of the 1960s believe that +each model was different and that these behaviors reflected microscopic differences. In addition, +Landau’s theory was regarded as phenomenological and believed to be of questionable validity. +3.3. The Renormalization Group +This situation was to change with the development of the Renormalization Group, due pri- +marily to the work of Leo P. Kadanoff [25, 26, 27] and Kenneth G. Wilson [28, 29, 30, 31, 32, 33]. +The renormalization Group was going to have (and still has) a profound effect both in Condensed +Matter Physics and in Quantum Field Theory (and beyond). +3.3.1. Scaling +Several phenomenological theories were proposed in the 1960s to describe the singular be- +havior of physical observables near a continuous phase transition [34, 35, 36]. These early works +argued that in order to explain the singular behavior of the observables the free energy density +had to have a singular part which should be a homogeneous function of the temperature, mag- +netic field, etc. A function f(x) is homogeneous if it satisfies the property that it transforms +irreducibly under dilations, i.e. f(λx) = λk f(x), there λ is a real positive number (a scale) and +k is called the degree. These heuristic ideas then implied that the critical exponents should obey +several identities. In 1966 Kadanoff wrote and insightful paper in which he showed that the ho- +mogeneity hypothesis implied that in that regime these systems should obey scaling. He showed +that this can be justified by performing a sequence of block-spin transformations in which the +configurations that vary rapidly at the lattice scale a become averaged at the scale of a larger sized +block of length scale ba > a which resulted in a renormalization of the coupling constants from +{K} at scale a to {K′} at the new scale ba [25, 27]. In other words, the block spin transformation +amounts to a scale transformation and a renormalization of the couplings (and operators). From +this condensed matter/statistical physics perspective the important physics is in the long distance +(“infrared”) behavior. +A significant consequence of these ideas was that close enough to a critical point, if the +distance |x − y| between two local observables O(x) and O(y) is large compared to the lattice +spacing a but small compared to the correlation length ξ, their correlation function takes the +form of a power law +⟨O(x)O(y)⟩ ∼ +const. +|x − y|2∆O +(3) +where ∆O is a positive real number known as the scaling dimension of the operator O [34, 25, 37]. +These conjectures were known to be satisfied in the non-trivial case of the 2D Ising model [26], +as well as in the Landau-Ginzburg theory once the effects of Gaussian fluctuations were included +[37]. For example, in the 2D Ising model, the scaling dimensions of the local magnetization σ is +∆σ = 1/8 and of the energy density ε is ∆ε = 1, which were sufficient to explain all the singular +behaviors known at that time. +The concept of renormalization actually originated earlier in quantum field theory as part +of the development of Quantum Electrodynamics (QED). In QED the notion of renormaliza- +tion was used to “hide” the short distance (“ultraviolet”) divergencies of the Feynman diagrams +needed to compute physical processes involving electrons (and positrons) and photons, i.e. their +strong, divergent, dependence of an artificially introduced short-distance cutoff or regulator. In +particular the sum of the leading diagrams that enter in the electron-photon vertex amounted to +5 + +a redefinition (renormalization) of the coupling constant. It was observed by Murray Gell-Mann +and Francis Low that this renormalization was equivalent to the solution of a first order differ- +ential equation that governed the infinitesimal change of the coupling, the fine structure constant +α = e2/4π, under an infinitesimal change of the UV cutoff Λ [38] +Λ dα +dΛ ≡ β(α) = 2 +3πα2 + O(α3) +(4) +where β(α) is the Gell-Mann-Low beta function. Except for the work by Nikolai Bogoliubov +and coworkers [39], this reinterpretation by Gell-Mann and Low was not actively pursued, partly +because it predicted that the renormalized coupling became very large at short distances, α → ∞, +and, conversely, it vanished in the deep long distance regime, α → 0 (if the electron bare mass is +zero). In other terms, QED is strongly coupled in the UV and trivial in the IR. The same behavior +was found in the case of the theory of a scalar field φ(x) with an φ4 interaction which is relevant +in the theory of phase transitions. In addition to these puzzles, the1960s saw the experimental +development of the physics of hadrons which involve strong interactions. For these reasons, for +much of that decade most high-energy theorists had largely abandoned the use of quantum field +theory, and explored other, phenomenologically motivated, approaches (which led to an early +version of string theory.) At any rate the notion that the physics may depend on the scale was +present as was the notion that in some regimes field theories may exhibit scale-invariance at least +in an approximate form. +3.3.2. The Operator Product Expansion +The next stage of the development of these ideas was the concept of the operator product +expansion (OPE). If we denote by {O j(x)} the set of all possible local operators in a theory (a field +theory or a statistical mechanical system near criticality), then the product of two observables on +this list closer to each other than to any other observable (and to the correlation length ξ) obeys +the expansion +lim +x→y O j(x)Ok(y) = lim +x→y +� +l +C jkl +|x − y|∆j+∆k−∆l Ol +� x + y +2 +� +(5) +where this equation should be understood as a weak identity, valid inside an expectation value. +Remarkably, this concept was derived independently and simultaneously by Leo Kadanoff [40] +(who was working in critical phenomena), by Kenneth Wilson [41] (who was interested in the +short distance singularities arising in Feynman diagrams), and by Alexander Polyakov [42, 43] +(also working in critical phenomena). In Eq.(5) {∆j} are the scaling dimensions of the operators +{O j}. The coefficients C jkl are (like the dimensions) universal numbers. In a follow up paper +Polyakov showed that if the theory has conformal invariance, i.e. scale invariance augmented by +conformal transformations which preserve angles, then, provided the operators O j are suitably +normalized, the coefficients C jkl of the OPE are determined by a three point correlator +⟨O j(x)Ok(y)Ol(z)⟩ = +C jkl +|x − y|∆jk|y − z|∆kl|z − x|∆l j +(6) +where ∆jk = ∆j + ∆k − ∆l. These results constitute the beginnings of Conformal Field Theory. +In a nontrivial check, Kadanoff and Ceva showed that the OPE holds for the local observables of +the 2D Ising model [44]. +3.3.3. Fixed Points +The next and crucial step in the development of the renormalization Group was made by +Kenneth Wilson. Wilson was a high-energy theorist who wanted to know how to properly define +a quantum field theory and the physical meaning of renormalization. +In a Lorentz invariant quantum field theory one is interested in the computation of the ex- +pectation value of time-ordered operators. In the case of a self-interacting scalar field φ(x) in +D-dimensional Euclidean space-time, obtained by analytic continuation from Minkowski space- +time to imaginary time, the observables are computed from the functional (or path) integral by +functional differentiation of the partition function +Z = +� +Dφ exp +� +−S (φ, ∂µφ) + +� +dDx J(x)φ(x) +� +(7) +6 + +with respect to the local sources J(x). For a scalar field the Euclidean action is +S = +� +dDx +�1 +2(∂µφ(x))2 + m2 +2 φ2(x) + λ +4!φ4(x) +� +(8) +which has the same form as the free energy of the Landau-Ginzburg theory of phase transitions +shown in Eq.(2). It is apparent that the Landau-Ginzburg theory is the classical limit of the theory +of the scalar field whose partition function is a sum over all histories of the field. It is easy to +see that en expansion of the partition function (or of a correlator) in powers of the the coupling +constant λ can be cast in the form of a sum of Feynman diagrams. To lowest order in λ a typical +Feynman diagram involves a one-loop integral in momentum space of the form +I(p) = +� +dDq +(2π)D +1 +(q2 + m2)((q − p)2 + m2) +(9) +As noted by Wilson in his Nobel Lecture [33], this integral has large contributions from the IR +region of small momenta q ∼ 0, but for any dimension D ≥ 4 has a much larger contribution +form large the UV region of large momenta, which requires the introduction of a UV cutoff Λ +in momentum space (or a lattice spacing a in real space by defining the theory on a hypercubic +lattice). In quantum field theory one then has to require that somehow one takes the limit a → 0 +(or Λ → ∞). To take this limit in the field theory is very much analogous to the definition of +a conventional integral in terms of a limit of a Riemann sum. The difference is that this is a +functional integral. While for a function of bounded variation in a finite interval (a, b) the limit +of a partition of the interval into N steps each of length ∆x, such that N∆x = b − a, exists and +defines the integral of the function +lim +∆x→0 lim +N→∞ +N +� +j=1 +f(x j)∆x j = +� b +a +dx f(x), +(10) +the analogous statement does not obviously exists in general for a functional integral, i.e. an +integral over a space of functions which is what is required. In fact, although thousands of +integrals of a function are known to exist, there are extremely few examples for a functional +integral. Moreover, in order to take the continuum limit the lattice spacing must approach zero, +a → 0. This means that physical scales, such as the correlation length ξ, must diverge in lattice +units so that they can be fixed in physical units. But to do that one has to be asymptotically +close to a continuous phase transition! Hence, the problem of defining a quantum field theory is +equivalent to the problem of critical phenomena at a continuous phase transition! +Wilson gave a systematic formulation to the Renormalization Group by generalizing the ear- +lier ideas introduced by Kadanoff and the earlier work by Gell-Mann and Low. Wilson’s key +contribution was the introduction of the concept of a fixed point of the Renormalization Group +transformation [28, 29, 30, 31]. As we saw, the block-spin transformation is a procedure for +coarse graining the degrees of freedom of a physical system resulting in a renormalization of the +coupling constants. Upon the repeated action of the RG transformation its effect can be pictured +as a flow in the space of coupling constants. However, in addition of integrating-out short dis- +tance degrees of freedom one needs to restore the units of length which have changed under that +process. This requires a rescaling of lengths. Once this is done, Wilson showed that the resulting +RG flows necessarily have fixed points, special values of the couplings which are invariant (fixed) +under the action of the RG transformation. He then deduced that at a fixed point the theory has +no scales, aside from the linear size L of the system and the microscopic UV cutoff (the lattice +spacing a in a spin system). +This analysis means that for length scales long compared to a → 0 but short compared to +L → ∞ the theory acquired a new, emergent, symmetry: scale invariance. Therefore, at a fixed +point the correlators of all local observables must be homogeneous functions (hence, must scale). +3.3.4. Universality +A crucial consequence of the concept if the fixed point is that phase transitions can be clas- +sified into universality classes. Universality means that a large class of physical systems with +different microscopic properties have fixed points with the same properties, i.e. the same scaling +dimensions, operator product expansions and correlation functions at long distances independent +on how they are defined microscopically. Although the renormalization group transformation is +a transformation scheme that we define and, because of that the location in coupling constant +7 + +space of the fixed point itself does depends on the scheme we choose, its universal properties are +the same. Thus, universality classes depend only on features such as the space (and spacetime) +dimension and the global symmetries of the system. But the systems themselves may be quite +different. This we speak if the Ising universality class in 2D, on the superfluid (or XY) transi- +tion class in 3D, etc. This concept, which originated in the theory of phase transition, has been +adopted and generalized in the development of conformal field theory. +3.3.5. RG flows +Combined with the condition that the correlators decay at long separations, homogeneity +implies that the correlators must have the form of Eq.(3). In addition, this equation also implies +that at a fixed point the operators (the local observables) have certain scaling dimensions. Let +us consider a theory close to a fixed point whose action we will denote by S ∗. Let {O j(x)} be +a complete set of local observables whose scaling dimensions are {∆j}. The total action of the +theory close to the fixed point then can be expanded as a linear combination of the operators with +dimensionless coupling constants {g j} +S = S ∗ + +� +dDx +� +j +g ja∆j−DO j(x) +(11) +Under a change of length scale x → x′ = bx, with b > 1, the operators (which must transform +homogeneously)change as O j(bx) = b−∆jO j(x). Since the phase space changes as dDx′ = bDddx, +we can keep the form of the action provided the coupling constants also change to compensate +for these changes as g′ +j = bD−∆jg j. Let b = |x′|/|x| = 1+da/a, where da is an infinitesimal change +of the UV cutoff a. Then, if we integrate-out the degrees of freedom in the range a < |x| < a+da, +the rate of change of the coupling constants {g j} under this rescaling is +adg j +da ≡ β(g j) = (D − ∆j)g j + . . . +(12) +which we recognize as a Gell-Mann Low beta function for each coupling constant. +This result says that if the scaling dimension ∆j < D, then the renormalized coupling will +increase as we increase the length scale, g′ +j > g j, and along this direction in coupling constant +space the RG flows away from the fixed point. Conversely, if ∆j > D the renormalized coupling +flows to smaller values, g′ +j < g j and the RG flows into the fixed point. We then say that an +operator is relevant if its scaling dimension satisfies ∆j < D, and that it is irrelevant if ∆j > D. +If ∆j = D then we say that operator is marginal. +To go beyond this simple dimensional analysis one has to include the effects of fluctuations. +To lowest orders in the couplings one finds [45] +adg j +da = (D − ∆j)g j + +� +k,l +C jkl gkgl + . . . +(13) +where {C jkl} are the coefficients of the OPE shown in Eq.(6). This expression is the general form +of a perturbative renormalization group and it is valid close enough to a fixed point. +Wilson and Fisher [30] used a similar approach to analyze how fluctuations alter the results +of the Landau-Ginzburg theory. They considered the partition function of Eq.(7) with the action +of Eq.(8). Instead of working in real space they considered the problem in momentum space and +partitioned the field configurations into slow and fast modes +φ(x) = φ<(x) + φ>(x) +(14) +where φ>(x) are configurations whose Fourier components have momenta in the range bΛ < |p| < +Λ, where Λ is a UV momentum cutoff and b < 1. Hence, if we choose b → 1, the fast modes φ > +(x) have components in a thin momentum shell near the UV cutoff Λ. Conversely, the slow modes +φ<(x) have momenta in the range 0 ≤ |p| < bΛ. One can then use Feynman diagrams to integrate +out the fast modes and derive an effective low-energy action with renormalized couplings. In the +case of a free field theory (with λ = 0) the scaling dimension of the φ4 operator is ∆4 = 2(D − 2) +whereas the φ2 operator has dimension ∆2 = D − 2. Upon defining a dimensionless mass and +coupling constant by m2 = tΛ2 and λ = gΛ4−D, the beta functions are found to be [10] +β(t) = − Λ dt +dΛ = 2t + g +2 − gt +2 + . . . +(15) +β(g) = − Λ dg +dΛ = (4 − D)g − 3 +2g2 + . . . +(16) +8 + +(where we absorbed an uninteresting factor if the definition of the coupling constant). +The RG flows of Eq.(16) show that the free-field (Gaussian) fixed point at g = 0 is stable for +D > 4 and the asymptotic IR behavior is the same as predicted by the Landau-Ginzburg theory. +However, for D < 4, the free-field fixed point becomes unstable and a new fixed point arises at +g∗ = 2 +3ǫ + O(ǫ2), where we have set ǫ = 4 − D. This is the Wilson-Fisher fixed point. At this +fixed point the correlation length diverges with an exponent ν = 1 +2 + ǫ +12 + O(ǫ2), which deviates +from the predictions of the Landau-Ginzburg theory. The small parameter of this expansion is ǫ, +and this is known as the ǫ expansion. +The Wilson-Fisher (WF) fixed point is an example of a non-trivial fixed point at which the +correlation length is divergent. It has only one relevant operator, the mass term, which in the IR +flows into the symmetric phase for t > 0 and flows to the broken symmetry for t < 0. Conversely, +in the UV it flows into the WF fixed point. For these reasons condensed matter physicists say +that this is an IR unstable (or critical) fixed point while high-energy physicists say that it is the +UV fixed point. At this fixed point a non-trivial field theory can be defined with non-trivial +interactions. UV fixed points also define examples of what in high-energy physics are called +renormalizable field theories and can be used to define a continuum field theory. +The D = 4-dimensional theory is special in that the φ4 operator is marginal. As can be seen +in Eq.(16), at D = 4 the beta function for the dimensionless coupling constant g does not have +a linear term and is quadratic in g. In this case the operator is marginally irrelevant, and its +beta function has the same behavior as the beta function of Gell-Mann and Low for QED. Such +theories are said to have a “triviality problem” since, up to logarithmic corrections to scaling, +there are no interactions in the IR and, conversely, become large in the UV. +There are also fixed points at which the correlation length ξ → 0. These fixed points are +IR stable (and in a sense trivial). These stable fixed points are sinks of the IR RG flows. Such +fixed points define stable phases of matter, e.g. the broken symmetry state, the symmetric (or +unbroken state), etc. However in the UV they are unstable and in high-energy physics such fixed +points correspond to non-renormalizable field theories. +More sophisticated methods are needed to go beyond the lowest order beta functions of +Eq.(13), and the computation of critical exponents beyond the leading non-trivial order. Pos- +sibly the record high-precision calculations have been done for φ4 theory for which the beta +function is know to O(ǫ5). This has been achieved using the method of dimensional regulariza- +tion [46, 47, 48] (with minimal subtraction). Special resummation methods (Borel-Pad´e) have +been used to do these calculations in D = 3 dimensions [49]. Remarkably, these results are +so precise that in the case of the superfluid transition, which is well described by a φ4 theory +with a complex field, the results could only be tested in the microgravity environment of the +International Space Station! +3.3.6. Asymptotic Freedom +There are several physical systems systems of great interest whose beta function has the form +β(g) = Ag2 + . . . +(17) +The coupling constant has a different interpretation in each theory and the constant A > 0, +opposite to the sign of the beta function found in QED and φ4 theory in D = 4 dimensions. +This beta function means that while the associated operator is marginal, with this sign is actually +marginally relevant. This also means that the fixed point is unstable in the IR but the departure +from the fixed point is logarithmically small. Conversely, the in the UV the RG flows into the +fixed point and the effective constant is weak at short distances. This is the origin of the term +asymptotic freedom [50]. The paradigmatic examples of theories with a beta function of this +form are the Kondo problem, the 2D non-linear sigma model, and the D = 4 dimensional Yang- +Mills non-abelian gauge theory. +The Kondo problem is the theory of a localized spin-1/2 degree of freedom in a metal. the +electrons of the metal couple to this quantum impurity through an exchange interaction of the +impurity and the magnetic moment density of the mobile electrons in the metal with coupling +constant J. This problem is actually one dimensional since only the s wave channel of the +mobile (conduction band) electrons actually couple to the localized impurity. In 1970 Philip +Anderson developed a theory of the Kondo problem in terms of the renormalization of the Kondo +coupling constant g as a function of the energy scale [51]. Anderson used perturbation theory +in J to progressively integrated out the modes of the conduction electrons close to an effective +9 + +bandwidth Ec and found that the beta function has the form of Eq.(17) +Ec +dJ +dEc += −ρJ2 + . . . +(18) +where ρ is the density of states at the Fermi energy of the conduction electrons. This work +implied that the free-impurity fixed point is IR unstable and that the effective coupling constant +J increases as the energy cutoff Ec is lowered. He argued that at some energy scale, the Kondo +scale, perturbation theory breaks down and that there is a crossover to a strong coupling regime +which is not accessible in perturbation theory. +Shortly thereafter, in 1973 Wilson developed a numerical renormalization group approach +which showed that the Kondo problem is indeed a crossover from the free impurity fixed point +to the “renormalized” Fermi liquid [32]. In addition, Wilson use the numerical renormalization +group to examine the approach to the strong coupling fixed point and showed that it is charac- +terized by a Wilson ratio, a universal number obtained from the low temperature specific heat +and the impurity magnetic susceptibility (in suitable units). Wilson’s numerical RG predicted +a number close to 2π for the this ratio. In 1980 N. Andrei and P. Wiegmann showed (indepen- +dently) that the Kondo problem is an example of an integrable field theory that can be solved +by the Bethe ansatz [52, 53]. Their exact result was consistent with Wilson’s RG, including the +numerical value of the Wilson ratio. +In 1972 Gerard ’t Hooft and Martinus Veltman showed that Yang-Mills gauge theory is renor- +malizable [46]. This groundbreaking result opened the door to use quantum field theory to de- +velop the theory of strong interactions in particle physics known as Quantum Chromodynamics +(QCD). In 1973 David Gross and Frank Wilczek [50] and, independently, David Politzer [54] +computed the renormalization group beta function of Yang-Mills theory with gauge group G and +found it to be of the same form as Eq.(17), +Λ dg +dΛ = − g3 +16π2 +11 +3 C2(G) + . . . +(19) +where g is the Yang-Mills coupling constant and Λ is a UV momentum scale. Here C2(G) is the +quadratic Casimir for a gauge group G. For S U(3), the case of physical interest, C2(S U(3)) = 3. +This result implies that under the RG at large momenta (short distances) the Yang-Mills coupling +constant flows to zero (up to logarithmic corrections). This result holds in the presence of quarks +provided the number of quark flavors is less than a critical value. Hence at short distances the +effective coupling is weak. Gross and Wilczek called this phenomenon asymptotic freedom. +This behavior was consistent with the observation of weakly coupled quarks in deep inelastic +scattering experiments. However, the flip side of asymptotic freedom is that at low energies +(long distances) the coupling constant grows without limit, which implies that at low energies +perturbation theory is not applicable. This strong infrared behavior suggested that in QCD quarks +are permanently confined in color neutral bound states (hadrons). However, unlike the Kondo +problem we just discussed, QCD is not an integrable theory (so far as we know) and to show that +it confines has required the development of Lattice Gauge Theory [55, 56]. To this date the best +evidence for quark confinement has been obtained using large-scale Monte Carlo simulations in +Lattice Gauge Theory [57]. +We close this subsection with a discussion of an important case: the non-linear sigma models. +The O(N) non-linear sigma model is the continuum limit of the classical Heisenberg model for a +spin with N components. Historically, the non-linear sigma model is the effective field theory for +pions in particle physics. We will discuss its role in the theory of quantum antiferromagnets in +subsection 4.3 and especially in the case of quantum antiferromagnetic spin chains in subsection +5.2. +The simplest non-linear sigma model is a theory of an N-component scalar field n(x) which +satisfies the unite length local constraint, n2(x) = 1. The Euclidean Lagrangian is +L = 1 +2g(∂µn(x))2 +(20) +where g is the coupling constant (the temperature in the classical Heisenberg model). At the +classical level, i.e. in the broken symmetry phase, where ⟨n⟩ � 0, this model describes the N − 1 +massless modes (Goldstone bosons) of the spontaneously broken O(N) symmetry. Dimensional +analysis shows that the coupling constant has units of ℓD−2. Hence, we expect to find marginal +behavior at D = 2. In 1975 Alexander Polyakov used a momentum shell renormalization group +10 + +in D = 2 dimensions and showed that the beta function of this model is (here a is the short- +distance cutoff) [58] +β(g) = adg +da = N − 2 +2π g2 + O(g3) +(21) +Hence, in D = 2 dimensions also this theory is asymptotically free. As in the other examples we +just discussed, asymptotic freedom here also implies that the coupling constant g grows to large +values in the low-energy (long-distance) regime. In close analogy with Yang-Mills theory in +D = 4 dimensions, Polyakov conjectured that the O(N) non-linera sigma model also undergoes +dynamical dimensional transmutation [50], that the global O(N) symmetry is restored and that +for all values of the coupling constant g the theory is in a massive with a finite correlation length +ξ ∼ exp((N − 2)/2πg). Extensive numerical simulations, used to construct a renormalization +group using Monte Carlo simulations [59], showed that there is indeed a smooth crossover from +the weak coupling (low temperature) regime to the high temperature regime where the correlation +length is finite. The non-linear sigma model is a renormalizable field theory in D = 2 dimensions +[60]. For D > 2 dimensions it can be studied using the 2 + ǫ expansion [60, 49], which predicts +the existence of a nontrivial UV fixed point and a phase transition from a Goldstone phase to a +symmetric phase. +It turns out that there is a significant number of asymptotically free non-linear sigma models +in D = 2 dimensions, many of physical interest [61], in particular non-linear sigma models +whose target manifold is a coset space, a quotient of a group G and a subgroup H. The O(N) non +linear sigma model is an example since the broken symmetry space leaves the O(N −1) subgroup +unbroken (the manifold of the Goldstone bosons). In that case the quotient is O(N)/O(N − 1) +which is isomorphic to the N −1 dimensional sphere S N−1. In later sections we will discuss other +examples in which more general non-linear sigma models play an important role. +Models on coset spaces arise in the theory of Anderson localization in D = 2 dimensions. +Anderson localization is the problem of a fermion (an electron) in a disordered system in which +the electron experiences a random electrostatic potential. In the limit of strong disorder Philip +Anderson showed that all one-particle states are exponentially localized and the diffusion con- +stant (and the conductivity) vanishes[62]. There was still the question of when it is possible +for the electron to have a finite diffusion constant (and conductivity). In D = 2 dimensions +the conductivity is a dimensionless number which suggests that this may be the critical dimen- +sion for diffusion. Abrahams, Anderson, Licciardello and Ramakrishnan used a weak disorder +calculation to construct a scaling theory that implied that in D = 2 dimensions the RG flow +of the conductivity at long distances (large samples) flows to zero and all states are localized +[63]. Shortly thereafter Wegner gave strong arguments that showed that the existence of diffu- +sion implied that there are low-energy “diffusson” modes which behaved as Goldstone modes of +a non-linear sigma model on the quotient manifold O(N+ + N−)/O(N+) × O(N−) in the “replica +limit” N± → 0 [64]. A field theory approach to this non-linear sigma model was developed by +McKane and Stone [65] and by Hikami [66]. +4. Quantum Criticality +Quantum criticality is the theory of a phase transition of a system (e.g. a magnetic system) +at zero temperature that occurs as a coupling constant (or parameter) is varied continuously. +Although not necessarily under that name, this question has existed as a conceptual problem for +a long time, In particular, already in 1973 Wilson considered the problem of the behavior of +quantum filed theories blow four spacetime dimensions and their phase transitions [67]. +The modern interest in condensed matter physics stems from discoveries made since the late +1980s. Since that time he behavior of condensed matter systems at a quantum critical point has +emerged as a major focus in the field. There were several motivations for this problem. One +was (and still is) to understand the behavior of quantum antiferromagnets in the presence of +frustrating interactions. Frustrating interactions are interactions which favor incompatible types +of antiferromagnetic orders. The result is the presence of intermediate non-magnetic “valence +bond” phases that favor the formation of spin singlets between nearby spins. these phases typ- +ically either break the point group symmetry of the lattice or are spin liquids (which will be +discussed below). Another motivation is that doped quantum antiferromagnets typically harbor +superconducting phases (among others) whose high-temperature behavior is a “strange” metal +that violates the basic assumptions (and behaviors) of Fermi liquids. The most studied version +of this problem is the case of the copper oxide high temperature superconductors. It was con- +jectured that there is a quantum critical point inside the superconducting phase at which the +11 + +antiferromagnetic order (or other orders) disappears and which may be the reason for the strange +metal behavior above the superconducting critical temperature. Many of these questions are +discussed in depth by Sondhi and coworkers [68] and in the textbook by Sachdev [69]. +4.1. Dynamic Scaling +We will consider a general quantum phase transition and assume that it is scale invariance. +However, except for the case of relativistic quantum field theories, in condensed matter systems +space and time do not need to scale in the same way. Let us assume that the system of interest has +just one coupling constant g and that the system of interest has a quantum phase transition (at zero +temperature) at some critical value gc between two phases, for instance one with a spontaneously +broken symmetry and a symmetric phase. If the quantum phase transition is continuous then the +correlation length ξ will diverge at gc and so will the correlation time ξt. However these two +scales are in general different and do not necessarily diverge at the same rate. So, in general, +if some physical quantity is measured at the quantum critical point at some momentum p and +frequency ω, the length scale of the measurement is 2π/|p| and at a frequency is ω ∼ |p|z, where +z is the dynamic critical exponent. Let us say that we measure the observable O at momentum p +and frequency ω at gc. Scale invariance in both space and time means that at gc the observable +O(p, ω) at momentum p and frequency ω must scale as +O(p, ω) = |p|−∆O ˜O(|p|z/ω) +(22) +where ∆O is the scaling dimension of the observable O. +The situation changes at finite temperature T. A quantum field theory at temperature T is +described by a path integral on a manifold which along the imaginary time direction τ is finite +of length 2π/T and periodic for a theory bosonic fields and anti-periodic for fermionic fields +[10]. Since the imaginary time direction is finite, the behavior for correlation times ξτ < 2π/T +and ξτ > 2π/T must be different. Indeed, in the first regime the behavior is essentially the same +as at T = 0, while in the second it should be given by the classical theory in the same space +dimension.At the quantum critical point gc there is only one time scale ξτ ∼ 2π/T and only one +length scale ξ ∼ (2π/T)1/z. +4.2. The Ising Model in a Transverse Field +The prototype of the quantum phase transition is the Ising model in a transverse field. This +model describes a system of spin-1/2 degrees of freedom with ferromagnetic interactions with +uniaxial anisotropy in the presence of a transverse uniform magnetic field. The Hamiltonian is +H = −J +� +⟨r,r′⟩ +σ3(r)σ3(r′) − h +� +r +σ1(r) +(23) +where J and h are the exchange coupling constant and the strength of the transverse field, re- +spectively. Here σ1 and σ3 are the two Pauli matrices defined on the sites {r} of a lattice with +ferromagnetic interactions between spins on nearest neighboring sites. The Hilbert space is the +tensor product of the states of the spins at each site of the lattice. At each site there are two nat- +ural bases of states: the eigenstates of σ3, which we denote by | ↑⟩ and | ↓⟩ (whose eigenvalues +are ±1), and the eigenstates of σ1, which we denote by |±⟩ (whose eigenvalues are also ±1). +It is well known that the Ising Model in a Transverse Field on a hypercubic lattice in D +dimensions is equivalent to a classical Ising model in D + 1 dimensions [70, 71]. These two +models are related through the transfer matrix. Indeed, a classical Ising model can be regarded +as a path integral representation of the quantum model in one dimension less. For simplicity +we will see how this work for the 2D the classical ferromagnetic Ising model of Eq.(1), but the +construction is general. We will regard the configuration of spins on a row of the 2D lattice as +a state of a quantum system, and the set of states on all rows as the evolution of the state along +the perpendicular direction that we will regard as a discretized imaginary time. The contribution +from two adjacent rows to the partition function defines the matrix element of a matrix between +two arbitrary configurations. In Statistical Physics this matrix is known as the Transfer Matrix ˆT +and the full partition function (with periodic boundary conditions) is +Z = tr ˆT Nτ +(24) +where Nτ is the number of rows. For the case of the ferromagnetic Ising model (actually, for any +unfrustrated model) the transfer matrix can always be constructed to be hermitian. This property +12 + +holds in fact for any theory that satisfies a property known as reflection positivity which requires +that all (suitably defined) correlation functions be positive. For theories of this type, and the Ising +model is an example, the matrix elements of the transfer matrix can be identified with the matrix +element of the evolution operator of a quantum theory for a small imaginary time step [70]. Also, +the positivity of the correlators is equivalent to the condition of positivity of the norm of states in +the quantum theory. +For the classical models that satisfy these properties, all directions of the lattice are equiv- +alent. Moreover, asymptotically close to the critical point, the behavior of all the correlators +becomes isotropic, i.e. invariant under the symmetries of Euclidean space. This means that +the arbitrary choice of the direction for the transfer matrix is irrelevant. Consequently, tin the +quantum model its equal-time correlators behave the same way as its correlation functions in +imaginary time. In other words space and time behave in the same way and the quantum the- +ory is relativistically invariant. This implies at the quantum critical point the energy ε(p) of its +massless excitations should behave as ε(p) = v |p|. In a relativistic theory the dynamical critical +exponent must be z = 1 and the coefficient v is the speed of the excitations (the “speed of light”). +We should note that this is not necessarily always the case. There are in fact classical systems, +e.g. liquid crystals [72], which are spatially anisotropic and map onto quantum mechanical theo- +ries in one less dimension for which the dynamical critical exponent z � 1. One such example are +the Lifshitz transitions of nematic liquid crystals in three dimensions and the associated quantum +Lifshitz model in D = 2 dimensions, for which the dynamical exponent is z = 2 [73]. +Just as in the classical counterpart in D + 1 dimensions, the quantum model in D ≥ 1 has +two phases: a broken symmetry ferromagnetic phase for J ≫ h and a symmetric paramagnetic +phase for h ≫ J. In the symmetric phase the ground state is unique (asymptotically is the +eigenstate of σ1 with eigenvalue +1), while in the broken symmetry phase the ground state +is doubly degenerate (and asymptotically is an eigenstate of σ3) and there is a non-vanishing +expectation value of the local order parameter ⟨σ3(r)⟩ � 0. In the symmetric phase the correlation +function of the local order parameter decays exponentially with distance with a correlation length +ξ, as does the connected correlation function in the broken symmetry phase. The model has +a continuous quantum phase transition at a critical value of the ratio h/J. For general space +dimensions D > 1 this model is not exactly solvable and much of what we know about it is due +to large-scale numerical simulations. +This problem was solved exactly in one-dimension [23] using the Jordan-Wigner transfor- +mation that maps a one dimensional quantum spin system to a theory of free fermions [22]. The +fermion operators at site j are +χ1(j) = K(j − 1)σ3(j), +χ2(j) = i ˆK(j)σ3(j) +(25) +where K(j) is the kink creation operator (i.e. the operator that creates a domain wall between +sites j and j + 1 [70], and is given by +K(j) = +� +n≤ j +σ1(n) +(26) +The operators χ1(j) and χ2(j) are hermitian, χ† +j(n) = χ j(n), and obey the anticommutation algebra +{χ j(n), χ j′(n′)} = 2δ j j′δnn′ +(27) +Hence, they are fermionic operators are hermitian, anti-commute with each other and square to +the identity. Operators of this type are called Majorana fermions. +Alternatively, we can use the more conventional (Dirac) fermion operators c(n) and its adjoint +c†(n) which are related to the Majorana fermions as +c(n) = χ1(n) + iχ2(n), +c†(n) = χ1(n) − iχ2(n) +(28) +which obey the standard anticommutation algebra +{c(n), c(n′)} = {c†(n), c†)n′)} = 0, +{c(n), c†(n′)} = δ − nn′ +(29) +In this sense, a Majorana fermion is half of a Dirac fermion. +In terms of the Majorana operators the Hamiltonian of Eq.(23) becomes +H = i +� +j +χ1(j)χ2(j) + ig +� +j +χ2(j)χ1(j + 1) +(30) +13 + +where we have rescaled the Hamiltonian by a factor of h and the coupling constant is g = J/h. +Here we have not specified the boundary conditions (which depend on the fermion parity). Qual- +itatively, the Majorana fermions can be identified with the domain walls of the classical models. +In the Ising model the number of domain walls on each row is not conserved but their parity is. +Likewise, the number of Majorana fermions NF is not conserved either but the fermion parity, +(−1)NF, is conserved. +It is an elementary excercise to show that the spectrum of this theory has a gap G(g) which +vanishes at gc = 1 as G(g) ∼ |g − gc|ν, with an exponent ν = 1. Since the Hamiltonian of +Eq.(30) is quadratic in the Majorana operators, these operators obey linear equations of motion. +In the scaling regime we take the limit of the lattice spacing a → 0 and the coupling constant +g → gc = 1, while keeping the quantity m = (g − gc)/a fixed. In this regime, the two-component +hermitian spinor field χ = (χ1, χ1), and χ† = χ, satisfies a Dirac equation +(i/∂ − m)χ = 0 +(31) +where we set the speed � = 1, and where defined the 2 × 2 Dirac gamma-matrices γ0 = σ2, +γ1 = iσ3, and γ5 = σ1. Upon defining ¯χ = χTγ0, we find that the Lagrangian of this field theory +is +L = ¯χi/∂χ − 1 +2m ¯χχ +(32) +which indeed becomes massless at the quantum phase transition of the Ising spin chain. For these +considerations, we say that the phase transition of the Ising model (2D classical or 1D quantum) +is in the universality class of massless Majorana fermions where m → 0. In Eq.(32) we have +used the standard Feynman slash notation, /a = γµaµ, where aµ is a vector. +4.3. Quantum Antiferromagnets and Non-Linear Sigma Models +As we noted above, the discovery of high temperature superconductors in the copper ox- +ide compounds prompted the study of the behavior of these strongly correlated materials at low +temperatures and of possible quantum phase transitions which they may host. The prototypical +cuprate material La2CuO4 is a quasi two-dimensional Mott insulator which exhibits long-range +antiferromagnetic order below a critical temperature Tc. A simple microscopic model is a spin-S +quantum Heisenberg antiferromagnet on the 2D square lattice of the Cu atoms, whose Hamilto- +nian is +H = 1 +2 +� +r,r′ +J(|r − r′|) S(r) · S(r′) +(33) +where S are the spin-S angular momentum operators. We will consider the case where the ex- +change interaction for nearest neighbors J is dominant and a weaker J′ ≪ J for next nearest +neighbors. In this section we do not consider the regime J′ ≃ J in which the interactions com- +pete for incompatible ground states due to frustration effects. +4.3.1. Spin coherent states +The simplest way to see the physics of this antiferromagnet is to construct a path-integral +representation for a spin-S system using spin coherent states [74, 75, 76]. For details see Ref.[9] +which we follow here. A coherent state of the (2S + 1-dimensional) spin-S representation of +SU(2) is the state |n⟩, labeled by the spin polarization unit vector n +|n⟩ = eiθ(n0×n)·S |S, S ⟩ +(34) +where n2 = 1. The states of the spin-S representation are spanned by the eigenstates of S 3 and +S2, +S 3 |S, M⟩ = M|S, M⟩, +S2|S, M⟩ = S (S + 1)|S, M⟩ +(35) +and |S, S ⟩ is the highest weight state with eigenvalues S and S (S + 1). In Eq.(34) n0 is a unit +vector along the axis of quantization (the direction e3), and θ is the colatitude, such that n · n0 = +cos θ. Two spin coherent states, |n1⟩ and |n2⟩, are not orthonormal, +⟨n1|n2⟩ = eiΦ(n1,n2,n0) S +�1 + n1 · n2) +2 +�S +(36) +where Φ(n1, n2, n0) is the area of the spherical triangle of the unit sphere spanned by the unit +vectors n1, n2 and n0. However, there is an ambiguity in the definition of the area of the spherical +14 + +triangle since the sphere is a 2-manifold without boundaries: if the “inside” triangle has spherical +area Φ, the complement (“outside”) triangle has area 4π − Φ. Thus, the ambiguity of the phase +prefactor of Eq.(36) is +ei4πS = 1 +(37) +since S is an integer or a half-integer. So, the quantization of the representations of SU(2) makes +the ambiguity unobservable. In addition, the spin coherent states |n⟩ satisfy the resolution of the +identity +I = +� +|n⟩⟨n| +�2S + 1 +4π +� +δ(n2 − 1) d3n +(38) +and +⟨n|S|n⟩ = S n +(39) +4.3.2. Path integral for a spin-S degree of freedom +As an example consider problem of a spin-S degree of freedom coupled to an external mag- +netic field B(t) that varies slowly in time. The (time-dependent) Hamiltonian is given by the +Zeeman coupling +H(t) = B(t) · S +(40) +As usual, the path-integral is obtained by inserting the (over-complete) set of coherent states at +a large number of intermediate times. The resulting path integral is a sum of the histories of the +spin polarization vector n(t) +Z = tr exp +� +i +� T +0 +dt H(t) +� += +� +Dn exp (iS[n]) +� +t +δ(n2(t) − 1) +(41) +where the action is +S = S SWZ[n] − S +� T +0 +dt B(t) · n(t) +(42) +where SWZ[n] is the Wess-Zumino action +SWZ[n] = +� T +0 +dt A[n] · ∂tn +(43) += +� 1 +0 +dτ +� T +0 +dt n(t, τ) · ∂tn(t, τ) × ∂τn(t, τ) +(44) +where A[n] is the vector potential of a Dirac magnetic monopole (of unit magnetic charge) at +the center of the unit sphere. The vector potential A[n] has a singularity associated with the +Dirac string of the monopole. We can write an equivalent expression which is singularity-free +using Stokes Theorem. We did this in the second line of Eq.(44) which required to extend the +circulation of A on the closed path described by n(t) to the flux of the vector potential through +the submanifold Σ of the unit sphere S 2 whose boundary is the history n(t), i.e. the area of Σ. +The smooth (and arbitrary) extension of configuration n(t) to the interior of Σ is done by defining +n(t, τ) such that n(t, 1) = n0, n(t, 0) = n(t), and n(0, τ) = n(T, τ). Since SWZ is the area of the +submanifold Σ of the unit sphere S 2, just as in Eq.(37), here too there is an ambiguity of 4π in +the definition of the area. Here too, this ambiguity is invisible since the spin S is an integer or a +half-integer. +The path integral of Eq.(41) was derived first by Michael Berry [77] (and extended by Barry +Simon [78]). The first term (which we called Wess-Zumino by analogy with its field theoretic +versions) is called the Berry Phase. The role of this term, which is first order in time derivatives, +is to govern the quantum dynamics of the spin which, in presence of a uniform magnetic field, +executes a precessional motion of the (Bloch) sphere. It is also apparent from this expression that +in the large-S limit, the path integral can be evaluated by means of a semiclassical approximation. +The coherent-state construction shows that this problem is equivalent to the path integral +of a formally massless non-relativistic particle of unit electric charge on the surface of the unit +sphere with a magnetic monopole of magnetic charge S in its interior! This is not surprising +since the Hilbert space of a non-relativistic particle moving on the surface of a sphere with and +radial magnetic field (the field of a magnetic monopole) has a Landau level type spectrum with +a degeneracy given by the flux. The condition of a massless particle means that only the lowest +Landau level survives and all other levels have an infinite energy gap. +The coherent state approach has been used to derive a path integral formulation for ferromag- +nets and antiferromagnets. A detailed derivation can be found in Ref. [9]. +15 + +4.3.3. Quantum Ferromagnet +We will consider first the simpler case of a quantum ferromagnet and in Eq.(33) we will set +J = −|J| < 0 for nearest neighbors and zero otherwise. The action for the path-integral for the +spin-S quantum Heisenberg ferromagnet on a hypercubic lattice is +S = S +� +r +SWZ[n(r, t)] − |J|S 2 +2 +� +⟨r,r′⟩ +� T +0 +dt �n(r, t) − n(r′, t)�2 +(45) +where we have subtracted the classical ground state energy. The oder parameter for this theory +is the expectation value of the local magnetization, n = ⟨n(r)⟩, which is constant in space but +points in an arbitrary direction in spin space. +In the low energy regime the important configurations are slowly varying in space and we +can simply approximate the action of Eq.(45) by its continuum version in d space dimensions +S = S +ad +0 +� +ddx SWZ[n(x, t)] − |J|S 2 +2ad +0 +� +ddx +� T +0 +dt (▽n(x, t))2 +(46) +where a0 is the lattice spacing. As before, the path integral; is done for a field which satisfies ev- +erywhere in space-time the constraint n2(x, t) = 1. This action can be regarded as non-relativistic +non-linear sigma model. +It is straightforward to show that the classical equations of motion for this theory are the +Landau-Lifshitz equations +∂tn = |J|S a2 +0 n × ▽2n +(47) +subject to the constraint n2 = 1. Due to the constraint, the Landau-Lifshitz equation is non-linear. +We will a decomposition of the field into a longitudinal and two transverse components, σ and +π, respectively +n = +�σ +π +� +(48) +subject to the constraint σ2 +π2 = 1. The linearized Landau-Lifshitz equations become (to linear +order in π) +∂tπ1 ≃ −|J|S a2 +0 ▽2 π2, +∂tπ2 ≃ +|J|S a2 +0 ▽2 π1 +(49) +The solution to these equations are ferromagnetic spin waves (magnons or Bloch waves) which +satisfy the dispersion relation +ω(p) ≃ |J|S a2 +0p2 + O(p4) +(50) +which shows that the dynamic exponent for a ferromagnet is z = 2. Notice that in this case the +two transverse components are not independent (they are effectively a dynamical pair). These +are the Goldstone bosons of a ferromagnet. +4.3.4. Quantum Antiferromagnet +Formally, the quantum antiferromagnet has a coherent state path integral whose action is +S = S +� +r +SWZ[n(r, t)] − JS 2 +2 +� +⟨r,r′⟩ +� T +0 +dt n(r, t) · n(r′, t) +(51) +with J > 0. For a bipartite lattice, e.g. the 1D chain, and the square and cubic lattices, the classi- +cal ground state is an antiferromagnet with a N´eel order parameter, the staggered magnetization. +Let m(r) be the expectation value of the local magnetization. A bipartite lattice is the union of +two interpenetrating sublattices, and the local magnetization is staggered, i.e. it takes values with +opposite signs (with equal values) on the two sublattices. Thus, we make the change of variables, +n(r, t) → (−1)rn(r, t) in Eq.(51) and find +S = S +� +r +(−1)rSWZ[n(r, t)] − JS 2 +2 +� +⟨r,r′⟩ +� T +0 +dt (n(r, t) − n(r′, t))2 +(52) +We want to obtain the low energy effective action for the field n(r, t). To this end, we decompose +this field into a slowly varying part, that we will call m(r, t), and a small rapidly varying part +l(r, t) (which represents ferromagnetic fluctuations) +n(r, t) = m(r, t) + (−1)ra0l(r, t) +(53) +16 + +Since n2(r, t) = 1, we will demand that the slowly varying part also obeys the constraint, +m2(r, t) = 1, and require that the two components be orthogonal to each other, m · l = 0. +Due to the behavior of the staggered Wess-Zumino terms of Eq.(52), the resulting continuum +field theory turns out to have subtle but important differences between one dimension and higher +dimensions. Here we will state the results for two and higher dimensions. We will discuss in +detail the one-dimensional below when we discuss the role of topology. +It turns out that if the dimension d > 1, the contribution of the staggered Wess-Zumino terms +for smooth field configurations is [75, 79, 80] +lim +a0→0 S +� +r +(−1)rSWZ[n(r, t)] = S +� +d3x l(x, t) · m(x, t) × ∂tm(x, t) +(54) +The continuum limit of the second term of Eq.(52) in the case of a two-dimensional system is +lim +a0→0 +JS 2 +2 +� +⟨r,r′⟩ +� T +0 +dt �n(r, t) − n(r′, t)�2 = a0 +JS 2 +2 +� +d3x +� +(▽m(x, t))2 + 4l2(x, t) +� +(55) +The massive field l[x, t] represents ferromagnetic fluctuations. Since this is a massive field it can +be integrated-out leading to an effective field theory for the antiferromagnetic fluctuations m(x, t) +whose Lagrangian is that of a non-linear sigma model +L = 1 +2g +� 1 +vs +(∂tm(x, t) − vs(▽m(x, t))2 +� +(56) +where the coupling constant is g = 2/S and the spin-wave velocity is vs = 4a0JS . If we to allow +for a weak next-nearest-neighbor interaction J′ > 0, the coupling constant g and the spin wave +velocity vs become renormalized to g′ ≃ g/ √1 − 2J′/J and v′ +s ≃ vs +√1 − 2J′/J. +We conclude that that the quantum fluctuations about a N´eel state are well described by a non- +linear sigma model. Provided the frustration effects of the next-nearest-neighbor interactions are +weak enough, the long-range antiferromagnetic N´eel order should extend up to a critical value +of the coupling constant gc where the RG beta function has a non-trivial zero, which signals a +quantum phase transition to a strong coupling phase without long-range antiferromagnetic order. +Motivated by the discovery of high temperature superconductivity in the strongly correlated +quantum antiferromagnet La2CuO4 (at finite hole doping) in 1988 Chakravarty, Halperin and +Nelson [81] utilized a quantum non-linear sigma model to analyze this system and its quantum +phase transition. La2CuO4 is a quasi-two-dimensional material and so it exhibits strong quantum +and thermal fluctuations. The upshot of this analysis is that while at T = 0 the non-linear sigma +model has a quantum phase transition, at T > 0 the long range order is absent in a strictly 2D +system but present in the actual material due to the weak-three-dimensional interaction. So, in +the strict 2D case there is no phase transition but two different crossover regimes: a renormal- +ized classical regime (without long range order), a quantum disordered regime and a quantum +critical regime. La2CuO4 has long range N´eel (antiferromagnetic) order at T = 0 and is in the +renormalized classical regime (with long range order due to the weak 3D interaction). +The non-linear sigma model does not describe the nature of the ground state for g > gc +beyond saying that there is no long range order. The problem is that, unlike the Ising model in a +transverse field, the microscopic tuning parameter is the next nearest neighbor antiferromagnetic +coupling J′, and to reach the regime g ≃ gc one has to make J′ ≃ J. This is the regime in which +frustration effects become strong. In this regime the assumption that the important configurations +are smooth and close to the classical N´eel state is incorrect. The nature of the ground state turns +out to depend on the value of S . +5. Topological Excitations +Topology has come to play a crucial role both in Condensed Matter Physics and in Quan- +tum Field Theory. Topological concepts have been used to classify topological excitations such +as vortices and dislocations and to provide a mechanism for phase transitions, quantum num- +ber fractionalization, tunneling processes in field theories, and nonperturbative construction of +vacuum states. Here we will discuss a few representative cases of what has become a very vast +subject. +17 + +5.1. Topological Excitations: Vortices and Magnetic Monopoles +In Condensed Matter Physics topological excitations play a central role in the description of +topological defects and on their role in phase transitions. Here topology integers in the classifi- +cation of the configuration space into equivalence classes characterized by topological invariants +[82]. The most studied example are vortices. Vortices play a key role in the mixed phase of type +II superconductors in a uniform magnetic field [83]. Vortices also play a key role in the Statistical +Mechanics of 2D superfluids and the the 2D classical XY model [84, 85] [86, 87] [88]. Disloca- +tions and disclinations play an analogous role in the theory of classical melting [84, 89, 90], and +2D and 3D classical liquid crystals [91, 92, 72]. +A similar problem occurs in Quantum Field Theory. Theories with global symmetries, such +as the two-dimensional O(3) non-linear sigma model discussed above, when formulated in Eu- +clidean space-time have instantons. Typically instantons are finite Euclidean action configu- +rations, which are also classified into equivalence classes (associated with homotopy groups) +labeled by topological invariants [93, 94]. Instantons play a central role in understanding the +non-perturbative structure of gauge theories. Gauge theories with a compact gauge group cou- +pled to matter fields have non-trivial vortex [95] and monopole [96, 97, 98] configurations, as +do non-abelian Yang-Mills gauge theories [99]. Instantons have also played a central role in +Condensed Matter Physics as well, notably in Haldane’s work on 1D quantum antiferromag- +nets (discussed below), and in the problem of macroscopic quantum tunneling and coherence +[100, 101]. +5.1.1. Vortices in two dimensions +In this section I will focus on the the problem of the superfluid transition in 2D and the +closely related problem of the phase transition of a magnet with an easy-plane anisotropy, the +classical XY model. A superfluid is described by an order parameter that is a one-component +complex field φ(x). If electromagnetic fluctuations are ignored, this description also applies +to a superconductor. The complex field can be written in terms of an amplitude |φ(x)|, whose +square represents the local superfluid density, and a phase θ(x) = arg(φ(x)). +Deep in the +super���uid phase the amplitude is essentially constant, that we will set to be a real positive +number φ0, while the phase field θ(x) is periodic with period 2π and can fluctuate. +Simi- +larly, an easy-plane ferromagnet is described by a two-component real order parameter field +M(x) = (M1(x), M2(x)) = |M(x)|(cosθ(x), sin θ(x)). Deep in the ferromagnetic phase the ampli- +tude |M| is essentially constant but the phase field θ(x) can fluctuate. +We will assume that we are in a regime where the local superfluid density |φ0|2 is well formed +(or, equivalently that |M| is locally well formed) but that the phase field is fluctuating. In this +regime the problem at hand is an O(2) ≃ U(1) non-linear sigma model, and its partition function +takes the form +Z = +� +Dθ exp +� +− +� +d2x 1 +2g +� +∂µθ(x) +�2� +(57) +where we defined the coupling constant g = T/J|φ0|2, where T is the temperature, J is an in- +teraction strength, and |φ0|2 is the magnitude (squared) of the amplitude of the order parameter, +which we will take to be constant; κ = J|φ0|2 is the phase stiffness. +Except for the requirement that the phase field be locally periodic, θ ≃ θ + 2π, superficially +this seems to be a trivial free (Gaussian) field theory. We will see that the periodicity (or, com- +pactification) condition makes this theory non-trivial. Indeed, configurations of the phase field +that are weak enough that that do not see the periodicity condition, for all practical purposes, can +regarded as being non-compact and ranging from −∞ to +∞. However there are many configu- +rations for which the periodicity condition is essential. Such configurations are called vortices. +Even in the absence of vortices, the periodic (compact) nature of the phase field is essential +to the physics of this problem. In fact the only allowed observables must be invariant under +local periodic shifts of the phase field. This implies that the phase field θ itself is not a physical +observable but that exponentials of the phase of the form exp(inθ(x)) are physical. This operator +is just the order parameter field of the XY model. In Conformal Field theory operators of this +type are called vertex operators [102, 103]. We will see below that this theory has a dual field ϑ, +associated with vortices, and that there are vertex operators of the dual field. In String Theory the +model of a compactified scalar is known as the compactified boson and represents the coordinate +of a string on a compactified space, in this case a circle S 1 [104]. +To picture a vortex consider a large closed curve C on the 2D plane. Hence, topologically +a closed curve is isomorphic to a circle, C ≃ S 1. The phase field θ(x) is equivalent to a unit +18 + +circle S 1. Therefore the configuration space are maps of S 1 (the large circle) onto S 1 (the unit +circle of the order parameter space). The configurations can be classified by the number of times +the phase winds on the large circle C. The winding number is an integer called the topological +charge of the configuration, the vorticity. Thus, a vortex is a configuration of the phase field θ(x) +that winds by 2πm (where m is an integer): +(∆θ)C +2π += 1 +2π +� +C +dx · ▽θ(x) ≡ i +� 2π +0 +dϕ +2π eiθ(ϕ)∂ϕe−iθ(ϕ) = m +(58) +where ϕ ∈ (0, 2π] is the azimuthal angle for a vector at the center of the large circle C. Here +n is the vorticity or winding number of the configuration; m > 0 is a vortex and m < 0 is an +anti-vortex. The vorticity is a topological invariant of the field the configuration θ(x) which does +not change under smooth changes. +The winding number of a vortex is a topological invariant that classifies the configurations +of the phase field as continuous maps of a large circle S 1 onto the unit circle S 1 defined by the +phase field. In Topology such continuous maps are called homotopies. The winding number +classifies these maps into a discrete set of equivalence classes, which form a homotopy group +under the composition of two configurations. In this case the homotopy group is called Π1(S 1). +Since the equivalence classes are classified by a topological invariant that takes integer values, +the homotopy group Π1(S 1) is isomorphic to the group of integers, Z [82]. +The field jµ(x) = ∂µθ(x) is the superfluid current, and the vorticity ω(x) is the curl of the +current, i.e. +ω(x) = ǫµν∂µ jν(x) = ǫµν∂µ∂νθ(x) +(59) +which vanishes unless θ(x) has a branch-cut singularity across which the phase field jumps by +2πn. Let ω(x) be the vorticity field with singularities at the locations {x j} of vortices with topo- +logical charge m j +ω(x) = 2π +� +j +m jδ2(x − x j) +(60) +which is satisfied by the phase field configuration +θ(x) = +� +j +2πm jIm ln(z − zj) +(61) +where we have used the complex coordinates z = x1 + ix2. Away from the singularities {x j}, +this configuration obeys the Laplace equation. Hence, it has a Cauchy-Riemann dual field ϑ(x) +which satisfies the Cauchy-Riemnann equation +∂µϑ = ǫµν∂νθ +(62) +which satisfies the Poisson equation +− ▽2ϑ(x) = ω(x) +(63) +whose solution is +ϑ(x) = +� +d2y G(|x − y|) ω(y) +(64) +where G(|x − y|) is the Green function of the 2D Laplacian +− ▽2G(|x − y|) = δ2(x − y) +(65) +In 2D this Green function is +G(|x − y|) = 1 +2π ln +� +a +|x − y| +� +(66) +where a is a short distance cutoff (a lattice spacing). In what follows we will assume that the +Green function of Eq.(66) has been cutoff so that G(|x − y|) = 0 for |x − y| ≤ a. +The energy of a configuration of vortices {n j} with vanishing total vorticity, � +j m j = 0, is +E[θ] = Jφ2 +0 +2 +� +d2x +� +∂µθ +�2 += Jφ2 +0 +2 +� +d2x +� +d2y ω(x) G(|x − y|) ω(y) = 2πJφ2 +0 +� +j>k +m jmk ln +� +a +|x j − xk +� +(67) +19 + +where we used that configurations with non-vanishing vorticity do not contribute to the partition +function since they have infinite energy in the thermodynamic limit. We conclude that, up to an +unimportant prefactor, that the partition function of Eq.(57) is the same as the partition function +a gas of charges {m j} (the vortices) with total vanishing vorticity, � +j m j = 0, +Z2DCG = +� +[{mj}] +exp +−2π Jφ2 +0 +T +� +j 2, relevant for ∆ < 2 +and marginal for ∆ = 2. Hence, vortices are marginal if ∆vortex = 2 which happens if g = π/2. +This is the same as to say that the system is at the Kosterlitz-Thouless critical temperature T = +TKT. Hence, vortices are irrelevant if T < TKT and relevant for T > TKT. +This RG analysis tells us that T > TKT, when vortices proliferate, the coupling constant ν +flows to strong coupling to a regime where ϑ is pinned to an integer value and the theory is in a +massive phase. In this phase the connected vortex correlator decays exponentially with distance +which is the same as to say that the vortex charge is screened. This is why in this phase the +vortices proliferate. It can be shown that in this phase the correlator of the spins of the XY +model, i.e. the correlator of the vertex operator exp(iθ(x)), decays exponentially with distance +and the systems is in its disordered phase. +There is still the question of the nature of the phase with T < TKT. The Mermin-Wagner The- +orem [107] (and its generalizations by Hohenberg [108] and Coleman [109]) states that classical +21 + +statistical mechanical systems with a global continuous symmetry group cannot undergo spon- +taneous symmetry breaking in space dimensions D ≤ 2 (and quantum systems with space-time +dimensions D ≤ 2). Does this theory violate this the Mermin-Wagner Theorem? The answer is +no. It is easy to see that in the phase in which the vortices are irrelevant, i.e. for T < TKT, the +correlator of the order parameter operator is always a power law of the distance |x − y|, +⟨exp(iθ(x)) exp(−iθ(y))⟩ = +const. +|x − y|g/2π +(81) +with an exponent that depends on temperature and satisfies +g +2π = +T +2πJφ2 +0 +≤ 1 +4 +(82) +In other words, the entire low temperature phase is not an ordered phase of matter since the cor- +relator is not constant at long distance. On the other hand, the correlators of the order parameter +exp(iθ) and of the vortices exp(iϑ) have a power la behavior for T < TKT, we conclude that in +this temperature range the system is scale invariant and that it has a line of critical points. +In summary, we succeeded in expressing the partition function as a sum over configurations +of the topological excitations, the vortices. We succeeded in doing that because vortices are +labeled by their coordinates on the plane. In addition, we found a non-trivial phase transition +since the entropy and the energy both scale logarithmically with the linear size of the system +or, equivalently, that vortices became marginally relevant at a critical temperature. This is the +mechanism behind the Kosterlitz-Thouless transition. +One may ask if this construction is generic and the answer is no. As an example consider the +Abelian Higgs model in D = 2. This model has a complex scalar field minimally coupled to a +Maxwell gauge field and, hence, its gauge group is U(1). In the classical spontaneously broken +phase, the gauge field becomes massive. In this phase the long range coherence of the phase field +of the vortices of the scalar field is screened at the scale of the penetration depth. As a result +the Euclidean action of the vortices is now finite. Furthermore, on longer scales the interaction +energy between vortices becomes short ranged. Hence, instead of a 2D Coulomb gas now one +has a gas of particles (vortices and anti-vortices) with short range interactions. In this case the +entropy always dominates and the vortices proliferate [110]. This behavior is quite analogous to +the restoration of symmetry by proliferation of domain walls in one-dimensional classical spin +chains [19] and to the analogous problem of tunneling in the path integral formulation of quantum +mechanical double-well potentials [93]. As a caveat, we should note that, in spite of the obvious +similarities, the 2D Abelian Higgs model does not describe correctly a 2D superconductor (i.e. a +superconducting film) since the electromagnetic field is not confined to the film and this renders +the electromagnetic action nonlocal. +5.1.2. Magnetic monopoles in compact electrodynamics +Instantons are of great interest in Quantum Field Theory since they provide for a mechanism +to understand the non-perturbative structure of these theories. For this reasons they have been +used to understand the mechanisms of quark confinement and the role of quantum anomalies +in non-Abelian gauge theories [99, 97, 98, 96, 111]. Instantons in non-Abelian gauge theories +are magnetic monopoles and the condensation of monopoles have long been argued to be the +mechanism behind quark confinement. +The simplest non-Abelian gauge theory that has magnetic monopoles is the Georgi-Glashow +[112]. This model has a three-component real field φ and an SU(2) Yang-Mills gauge field Aµ. In +its Higgs phase the scalar field φ acquires an expectation value which breaks the gauge symmetry +group SU(2) down to its diagonal U(1) subgroup. Since U(1) ⊂ SU(2), this Abelian gauge group +is compact, meaning that its magnetic fluxes are quantized. Polyakov [113] and ’t Hooft [111] +showed that in 2+1 Euclidean dimensions have non-singular instanton solutions which at long +distances resemble the magnetic monopole originally proposed by Dirac in 1931 [114] +Bi(x) = q +2 +xi +|x|2 − 2πqδi,3δ(x1)δ(x2)θ(−x3) +(83) +The first term in Eq.(83) is the magnetic radial field of a monopole of magnetic charge q. The +second term represents an infinitely long infinitesimally thin solenoid ending at the location of +the monopole, x = 0, that supplies the quantized magnetic flux 2πq. This singular term is +known as the Dirac string. The string itself (and its orientation) is physically unobservable to any +22 + +electrically charged particle that obeys the Dirac quantization condition, qe = 2π (in units where +ℏ = c = 1). +In the language of a lattice gauge theory [32], a theory with a compact (i.e. periodic) U(1) +gauge fields on a D = 3 cubic lattice, describing a compact gauge field in 2 + 1 dimensions. +This theory should have instantons that resemble magnetic monopoles much in the same way as +a theory with a compact global U(1) symmetry has vortices. The simplest example is Polyakov’s +compact electrodynamics [97] whose partition function is +Z = +� +x,µ +� 2π +0 +dAµ(x) +2π +exp + +1 +4e2 +� +x,µ,ν +cos(Fµν(x)) + +(84) +where Fµν(x) = ∆µAν(x)−∆νAµ(x) ≡ � +µ Aµ is the magnetic flux through the elementary plaquette +labeled by a site x and a pair of directions, µ and ν, with µ = 1, 2, 3. This theory is invariant under +local gauge transformations Aµ(x) → Aµ(x) + ∆µΦ(x) and it is also invariant under local periodic +shifts of the gauge fields Aµ(x) → Aµ(x) + 2πℓµ(x), where ℓµ(x) ∈ Z. The plaquette flux operator +satisfies the lattice version of the Bianchi identity that the product of exponentials of the flux on +the faces of every elementary cube of the lattice is +� +cubefaces +eiFµν(x) = 1 +(85) +which says that the theory can have magnetic monopoles of integer magnetic charge. +We will analyze this theory following an approach analogous to what we used for vortices in +section 5.1.1.To this end we will consider the partition function +Z[Bµν] = +� +DAµ exp +� +− 1 +4e2 +� +d3x +� +Fµν(x) − Bµν(x) +�2� +(86) +where Fµν(x) = ∂µAν − ∂νAµ is the field strength of the abelian U(1) Maxwell gauge field Aµ. +Here Bµν(x) is an (anti-symmetric) two-form background gauge field. The coupling constant of +this theory is e2. Since Aµ is a connection it has units of length−1, and F2 +µν is a dimension 4 field. +Then, in D = 3 dimensions, e2 has units of length−1. +The theory is invariant under two local transformations, namely the usual invariance under +gauge transformations +Aµ(x) → Aµ(x) + ∂µΦ(x), +Bµν(x) → Bµν(x) +(87) +where Φ(x) is an arbitrary smooth function of x. The presence of the background two-form field +Bµν now requires invariance under one-form gauge transformations +Aµ(x) → Aµ(x) + aµ(x), +Bµν(x) → Bµν(x) + ∂µaν − ∂νaµ +(88) +The two-form gauge field Bµν essentially represents the magnetic monopoles. Let {m j} be a +configuration of monopoles of charges m j with coordinates {x j}, with total vanishing monopole +charge, � +j m j = 0. Let M(x) be the magnetic monopole density at x, +M(x) = 2π +� +j +m j δ3(x − x j) +(89) +which can be expressed as the curl of the two-form gauge field Bµν, +M(x) = 1 +2ǫµνλ∂µBνλ(x) +(90) +We will proceed next much in the same way as in Eq.(72) and rewrite the partition function +of Eq.(86) in terms of a two-form Hubbard-Stratonovich field bµν(x) such that +Z[B] = +� +DAµ +� +Dbµν exp +� +−e2 +4 +� +d3x b2 +µν(x) + i +� +d3x 1 +2bµν(x) +� +Fµν(x) − Bµν(x) +�� += +� +DAµ +� +Dbµν exp +� +−e2 +4 +� +d3x b2 +µν(x) + i +� +d3x +� +Aµ(x)∂νbµν(x) − 1 +2bµν(x)Bµν(x) +�� +(91) +23 + +Thus, the gauge field Aµ plays the role of a Lagrange multiplier field the enforces the constraint +∂νbµν(x) = 0 +(92) +which is solved in terms of a compact scalar field ϑ(x) +bµν(x) = ǫµνλ∂λϑ(x) +(93) +Using this identity and the definition of the monopole density M(x) we find that the partition +function Z[Bµν] of Eq.(86) becomes +Z[B] = +� +Dϑ exp +� +− +� +d3x +�e2 +2 (∂µϑ(x))2 + iM(x)ϑ(x) +�� += +� +Dϑ exp +−e2 +2 +� +d3x(∂µϑ(x))2 + 2πi +� +j +m jϑ(x j) + +(94) +which requires that the field ϑ obeys the compactification condition ϑ → ϑ + n, where n is an +arbitrary integer. Eq.(94) says that the magnetic monopole instantons of the compact U(1) gauge +theory are dual to charges of the dual phase field ϑ, which has a compact U(1) global symmetry. +The full partition function is obtained by summing over all monopole configurations satis- +fying the total neutrality condition, � +j m j = 0. As in section section 5.1.1, we will weigh the +configurations with a coupling u and find +Z = +� +{mj} +Z{m j} +� +Dϑ exp +−e2 +2 +� +d3x (∂µϑ)2 + +� +j +2πim jϑ(x j) − u +� +j +m2 +j + +(95) +which is the same theory we found in Eq.(77) except that now we are in 3D. Moreover, summing +only over dilute configurations of monopoles and anti-monopoles we find, once again the sine- +Gordon theory but now in D = 3 dimensions: +Z = +� +Dϑ exp +� +− +� +d3x +�e2 +2 (∂µϑ)2 − � cos(2πϑ) +�� +(96) +with � = 2 exp(−u)/a3. +In spite of the similarities between Eq.(96) and the sine-Gordon theory in 2D, Eq.(78), the +physics is very different. It is straightforward to see that, in the limit v = 0, the monopole operator +correlator is +⟨exp �2πiϑ(x)) exp(−2πiϑ(y)� = exp +�4π2 +e2 [G(|x − y|) − G(0)] +� +≃ exp +� π +2e2 +� 1 +R − 1 +a +�� +(97) +where a is the short-distance cutoff. Unlike the behavior of the correlator of the vortex operators +in 2D found in Eq.(79), Eq.(97) does not show a power-law behavior. The reason is that at � = 0 +the compactified field ϑ should be regarded as the Goldstone boson of a spontaneously broken +U(1) symmetry. However, the cosine operator is now always relevant and the field ϑ is pinned +and its fluctuations are actually massive. +Looking back at the partition function of Eq.(95), we could integrate out the field ϑ and +obtain an expression with the same form as the Coulomb gas of Eq.(68) except that now this is +the three-dimensional neutral Coulomb gas [97] +Z3DCG = +� +[{mj}] +exp +− π +2e2 +� +j 0, one can always +rescale the time and space coordinates without affecting the form of the Lagrangian, including +the coupling constant or, as we will see, the value of the θ angle. In what follows we will assume +that we have done the rescaling in such a way that we set vs = 1 and, that time and space scale as +lengths. Thus we will use a relativistic notation and, after an analytic continuation to imaginary +time, we label the time coordinate by x2 and the space direction by x1. The partition function +now takes the form +Z = +� +Dmexp +� +− 1 +2g +� +Ω +d2x +� +∂µn +�2 + iθ Q[m] +� +(102) +where Ω is the spacetime manifold. In this notation the first term of the exponent is called the +Euclidean action of the non-linear sigma model. In Eq.(102) we denoted by Q[m] the quantity +Q[m] = 1 +8π +� +Ω +d2x ǫµνm · ∂µm × ∂νm +(103) +Here we will consider the case in which the spacetime manifold Ω is closed. In particular +we will assume that it is a two-sphere S 2. The quantity Q[m] is the integral of a total derivative +which counts the number of times the field configuration m(x1, x2) wraps around the sphere S 2. +In other words, it yields a non-vanishing result only for “large” configurations which wind (or +wrap) around the sphere S 2. Since Q[n] is an integer, it has the same for all smooth the field +configuration m that can be smoothly deformed into each other and are homotopically equiva- +lent. If we demand that the field configurations m have finite Euclidean action, which requires +that at infinity the configurations take the same (but arbitrary) value of m, we have effectively +compactified the x1 − x2 plane into a two sphere S 2. On the other hand the field m is restricted +by the constraint m2 = 1 to take values on a two-sphere S 2. Therefore, the field configurations +m(x1, x2) are smooth maps of the S 2 of the coordinate space to the S 2 of the target space of the +field m. Hence, the integer Q[m] classifies the smooth field configurations into a set of equiva- +lence classes each labeled by the integer Q. Under composition homotopies form groups, and the +equivalence classes themselves also form a group which, in this case, is isomorphic to the group +of integers, Z. In Topology, these statements are summarized by the notation Π2(S 2) ≃ Z. The +25 + +configurations with non-zero values of Q are called instantons which, in the quantum problem, +represent tunneling processes of the non-linear sigma model. +Another consequence of Q being an integer is that the contribution of its term to the weight of +the path integral of Eq.(103) is a periodic function of θ angle. On the other hand, since the only +allowed values of the θ are θ = 0 (mod 2π) for S integer, or θ = π (mod 2π) for S a half-integer, +the contribution of the topological invariant Q to the weight of the path integral is +exp(iθ Q[m]) = (−1)2S Q +(104) +Therefore, for spin chains with S integer, the weight is 1 and the topological invariant does not +contribute to the path integral. But, if S is a half-integer, the weight is (−1)Q[m], and it does +contribute. Moreover, its contribution is the same for all half-integer values of S . +These results have important consequences for the physics of spin chains which led Haldane +to some startling conclusions [115]. In the weak coupling regime, g ≪ 1 (equivalently, for +large S ), we can use the perturbative renormalization group and derive the beta function for all +these O(3) non-linear sigma models (with or without topological terms) and find that their beta +functions are the same as in Eq.(21) with N = 3. Hence, for all S , the effective coupling flows +to large values. We can make this inference since the topological term yields no contribution for +all configurations which are related by smooth deformations. Thus, we infer that all spin chains +with S integer are in a massive phase with an exponentially small energy gap ∼ exp(−2πS ). This +result is nowadays known as the Haldane gap. +On the other hand, these results also imply that all spin chain with half-integer spin S are +also the same and, in particular, th same as spin-1/2 chains. However, the Hamiltonian of the +quantum spin chain with spin-1/2 degrees of freedom is an example of an integrable system and +its spectrum is known to be gapless from its Bethe Ansatz solution [116, 117]. However, the +spin-1/2 chain is not only gapless but its low energy states are gapless solitons with a relativistic +spectrum. The low energy description of a theory of this type must be described by a conformal +field theory. Haldane concluded that the RG flow for spin-1/2 chains must have an IR stable fixed +point at some finite (and large) value of the coupling constant. This conjecture was confirmed by +Affleck and Haldane [118] who showed that the spin-1/2 Heisenberg chains are in the universality +class of the SU(2)1 Wess-Zumino-Witten model [119] whose CFT was solved by Knizhnik and +Zamolodchikov [120]. +5.3. Topology and open integer-spin chains +In the preceding section we showed that integer spin chains have a Haldane gap. We did that +by showing that in that case the topological term is absent. However, the derivation is correct +provided the spacetime manifold is closed, e.g. a sphere, a torus, etc. What happens if the system +has a boundary? Let us denote the boundary of Ω by Γ = ∂Ω. For example we will take Γ to be +along the imaginary time direction and hence that it is a circle of circumference 1/T, where T +is the temperature. The topological term has the same form as the Berry phase term of the path +integral for spin, the “Wess-Zumino” term of Eq.(44), but its prefactor is 1/2 as big. Thus, for +a system with an open boundary the topological term yields a net contribution equal to a Berry +phase with a net prefactor of S/2. +In other words, the boundary of the integer spin chain behaves as a localized degree of free- +dom whose spin is 1/2 (mod an integer). Form the periodicity requirement we also see that if +the spin chain is made of odd-integer degrees of freedom, there should be a spin 1/2 degree of +freedom localized at the open boundary!. Conversely, if the chain is made of even-integer spins, +there is no boundary degree of freedom! This line of argument implies that an antiferromagnetic +chain with odd-integer spins must have a spin-1/2 degree of freedom at the boundary whereas +a chain of even-integer spins should not. Notice that the “bulk” behavior is the same for both +odd and even integer spin chains. The difference is whether or not they have a non-trivial “zero- +mode” state at the boundary. In particular, the existence of this state is robust, i,.e. it cannot +be removed by making smooth changes to the quantum Hamiltonian or, what is the same, the +boundary state is topologically protected. +A simple system that displays a protected spin-1/2 zero mode at the boundary is a generalized +S = 1 spin chain with Hamiltonian +H = α +N +� +j=1 +S(j) · S(j + 1) + β +N +� +j=1 +(S(j) · S(j + 1))2 +(105) +26 + +where S(j) = (S x(j), S y(j), S z(j) are the spin 1 matrices at each lattice site j. This problem was +examined in great detail by Affleck, Kennedy, Lieb and Tasaki [121] who showed that at the +special value of the parameters α = 1/2 and β = 1/6 this Hamiltonian takes the form of a sum of +projection operators +H = +� +j +P2(S(j) + S(j + 1)) +(106) +where P2 is an operator that projects out the spin 2 states. These authors constructed the exact +ground state, known as the AKLT state, of this Hamiltonian by writing each spin 1 degree of +freedom of two spin-1/2 degrees of freedom at each site. They showed that the ground state is a +projected product state in which the “constituent’ spin-1/2 degrees of freedom (each labeled by ++ and − respectively) on nearby sites j and j + 1 are in a valence bond singlet state of the form +1√ +2(| ↑j,+, ↓j+1,−⟩− ↓j,+, ↑j+1,−⟩ (and symmetrizing at each site to project onto a spin 1 state). +This state of the spin-1 chain is translation invariant and gapped and hence agrees with Haldane’s +result. Moreover, it has a spin-1/2 degree of freedom at each open boundary [122]. +The arguments discussed above show that the AKLT state is a gapped topological state. The +gapless spin-1/2 boundary states are an example of spin fractionalization. The spin-1/2 boundary +degrees of freedom are an example of an edge state which are present in many, thought not all, +topological phases of matter. One may ask what symmetry protects the gaplessness of these edge +degrees of freedom. The only perturbation that would give a finite energy gap to the spin-1/2 +edge states is an external magnetic ���eld. However, this perturbation would break the global +SU(2) symmetry of the Hamiltonian as well as time reversal invariance. +6. Duality in Ising Models +Duality plays a significant role of our understanding of statistical physics and of quantum +field theory. Many seemingly unrelated correspondences between different theories have come +to be called dualities. +6.1. Duality in the 2D Ising Model +One of the earliest versions of duality transformations was used to relate the high-temperature +expansion of the 2D classical Ising model and its low-temperature expansion [123]. In all dimen- +sions, for concreteness we will think of a hypercubic lattice, the high temperature expansion is a +representation of the partition function as a sum of contributions of closed loops on the lattice. At +temperature T, a configuration of loops γ contributes with a weight which in the Ising model has +the form C(γ) tanhL(γ)(β), where β = 1/T and L(γ) is the length of the loop γ, i.e. the number +of links on the loop, and C(γ) is an entropic factor that counts the number of allowed loops with +fixed perimeter L(γ). However not all loop configurations are allowed as in the Ising model they +satisfy constraints such as being non-overlapping, etc. +In section 4.2 we noted that the partition function of the classical Ising model (in any dimen- +sion) can be interpreted as the path-integral of a quantum spin model in one dimension less on a +lattice with a discretized imaginary time. In this picture, the loops γ can be regarded as processes +in which pairs of particles are created at some initial (imaginary) time, evolve and eventually are +annihilated at a later (imaginary) time. In other words, the high temperature phase is a theory of +a massive scalar field. The restrictions on the allowed loop configurations represents interactions +among these particles. In the temperature range in which the expansion in loops is convergent +the particles are massive as the loops are small. As the radius of convergence of the expansion is +approached, longer and increasingly fractal-like loops begin to dominate the partition function, +and concomitantly the mass of the particles decreases. This process is the signal of the approach +to a continuous phase transition where the particles become massless. In fact, right at the critical +point the particle interpretation is lost as the associated fields acquire anomalous dimensions. +Returning to the 2D classical Ising model, Kramers and Wannier also considered the low +temperature expansion. This is an expansion around one of the broken symmetry state, e.g. +the state with all spins up. In this low temperature regime the partition function is a sum of +configurations of flipped spins. A typical configuration is a set of clusters of flipped spins. In the +absence of a uniform field, a configuration of flipped spins has a energy cost only on links of the +lattice with oppositely aligned spins (“broken bonds”). Thus a cluster of flipped spins costs an +energy equal to 2 (I assumed that I set J = 1) for each broken bond at the boundary of the cluster. +This boundary is domain wall which is a closed loop on the dual lattice. In 2D the dual of the +square lattice is the square lattice of the dual sites (the centers of the elementary plaquettes of the +27 + +2D lattice). In 2D the links of the direct lattice pierce the links of the dual lattice. We can see +that this analogous to the the geometric duality of forms: sites (“0-forms”) are dual to plaquettes +(“2-forms”) and links (“1-forms”) are dual to links (also “1-forms”). +Thus, in 2D the low temperature expansion is an expansion in the loops γ∗ of the dual lattice +that represent the domain walls. We can also regard the domain walls (the dual loops) as the +histories of of pairs of particles on he dual lattice. However, the weight of each dual loop is +exp(−2β) per link of γ∗. Except for that, the counting and restrictions on the dual loops γ∗ +are the same as those of γ. This means that there is a one-to-one correspondence between the +two expansions with the replacement tanh β ↔ exp(−2β∗). This mapping means that the dual +of the 2D classical Ising model (with global Z2 symmetry) at inverse temperature β is a dual +Ising model (also with global Z2 symmetry) on the dual lattice at inverse temperature β∗. This +correspondence is in close analogy with what we discussed in section 5.1.1. +In particular if one assumes that there is a transition at βc = 1/Tc then there should also be a +transition at β∗ +c = −1/2 lntanh βc. Moreover, if one further assumes (as Kramers and Wannier did +[123]) that there is a unique transition (correct in the Ising model but not in other cases), then the +critical point must be such that exp(−2βc) = tanh βc, which yields the value Tc = 2/ ln( +√ +2 + 1), +which agrees with the Onsager result [20]. +In section 4.2 we showed that the classical 2D Ising model is equivalent to the one-dimensional +Ising model in a transverse field whose Hamiltonian is given in Eq.(23). The 1D Ising model on +a transverse field has spin degrees of freedom defined on the sites of a one-dimensional chain +labeled by an integer-valued variable n. The 1D Hamiltonian is expressed in terms of local op- +erators, the Pauli matrices σ3(n) and σ1(n). The 1D chain has a dual lattice whose sites are the +midpoints of the chain. Thus in 1D sites (“0-forms”) are dual to links (“1-forms”) and viceversa. +We will now see that there is a Hamiltonian version of the Kramers-Wannier duality [70]. +In Eq.(26) we introduced the kink creation operator which flips are σ3 operators to the left and +including site j. For clarity we will denote the kink creation operator operator as τ3(˜n), where +˜n is the site of the dual lattice between the sites n and n + 1 of the original lattice. We will now +define an operator τ1(˜n) = σ3(n)σ3(n+1). The operators τ1(˜n) and τ3(˜n) satisfy the same algebra +os the Pauli operators σ1(n) and σ3(n). Furthermore, we readily find the Hamiltonian of the dual +theory is the same (up to boundary conditions) as the original Hamiltonian of Eq.(23) except +that the dual coupling constant is λ∗ = 1/λ. So, once again, if we assume that there is a single +(quantum) phase transition we require λ∗ +c = λc which is only satisfied by λc = 1. In this language +the kink creation operator plays the role of a disorder operator [70]. +6.2. The 3D duality: Z2 gauge theory +We will now discuss the role of duality in the 3D Ising model on a cubic lattice. This case, +and its generalizations to higher dimensions, was considered first by Franz Wegner [124]. +In section 6.1 we discussed the loop representation of the high temperature expansion and +showed that it has the form form in all dimensions. Hence, in 3D the high temperature expansion +is an expansion in closed loops γ with weight of tanh β per unit link of loop. Hence, in 3D as +well, the high temperature phase can be regarded as field theory of a massive scalar field. As in +the 2D case, the weight of a loop configuration is tanh β for each link of the loop γ. +However, the low temperature expansion has a radically different form and physical inter- +pretation. Much as in 2D, the low temperature expansion is an expansion in clusters of flipped +spins. Here too, in the absence of an external field, the only energy cost resides at the boundary +of the clusters of overturned spins, the domain walls. But in 3D the clusters occupy volumes +whose boundaries are closed surfaces Σ∗. In 3D a link (bond) is dual to a plaquette (expressing +the fact that in 3D a 1-form is dual to a 2-form). Hence the closed domain walls of overturned +spins is dual to a closed surface Σ∗ on the dual lattice. The weight of each configuration of closed +surfaces is exp(−2β) for each plaquette of the surface Σ∗. +These facts mean that the dual of the 3D Ising model is a theory on the dual lattice with +coupling constant β∗, with exp(−2β∗) = tanh β, such that its expansion for small β∗ is a sum over +configurations of closed surfaces σ∗, and for large β∗ is a sum of configurations of closed loops +γ∗ on the dual lattice. The dual theory is the naturally defined on (dual) plaquettes, not on links. +To this end, let us define a set of Ising-like degrees of freedom σµ(x) = ±1 located on the links +(x, µ) of the dual lattice. These degrees of freedom are coupled on each plaquette of the lattice. +Since each plaquette has four links, the coupling involves the degrees of freedom on all four links +28 + +of each plaquette. The partition function of the dual theory is +Z = +� +{σµ(x)} +exp +β∗ +� +plaquettes +σµ(x)σν(x + eµ)σµ(x + eµ + eν)σν(x) + +(107) +where the sum in th exponent (the negative of the “action”) runs on all the plaquettes of the dual +lattice. +The action of the theory of Eq.(107) is invariant under the reversal of all six Ising degrees of +freedom on links sharing a given site x. This is a local symmetry. Unlike the 3D Ising model, +which has a global Z2 symmetry of flipping all spins simultaneously, this theory has local (or +gauge) Z2 symmetry. This is the simplest example of a lattice gauge theory in which the degrees +of freedom are gauge fields that take values on the Z2 gauge group [124, 32, 71]. +We saw that in the spin model the high temperature expansion (i.e. the expansion in powers of +tanh β) is a sum over loop configurations which can be interpreted in terms of processes in which +pairs of particles are created, evolve (in imaginary time) and then are destroyed (also in pairs). +The analogous interpretation of the expansion of the Z2 gauge theory in powers of tanh β∗ = +exp(−2β) as a sum over the configurations of closed surfaces is not in terms of the histories of +particles but in terms of the histories of closed strings which, as they evolve, sweep the closed +surfaces. The physics is, however, more complex as the sum over surfaces runs over all surfaces +of arbitrary topology with arbitrary number of handles (or genus). Thus, over (imaginary) time a +closed strong is created, evolves, splits into two closed strings, etc. +Therefore, the 3D Ising model, which has a Z2 global symmetry is dual to a Z2 gauge theory +which has a Z2 local symmetry. The duality maps the high temperature (disordered) phase of +the Ising model to the strong coupling (small β∗) phase of the gauge theory. In the gauge theory +language this is the confining phase. This can be seen by computing the expectation value of the +Wilson loop operator on the closed loop Γ, which here reads ⟨� +(x,µ)∈Γ σµ(x)⟩. In the small β∗ +phase this expectation value decays exponentially with the size of the minimal surface bounded +by the loop Γ. This behavior of the area law of the Wilson loop. Under duality the insertion of +this operator is equivalent to an Ising model with a domain wall terminating on the loop Γ, and +the area law of the Wilson loop is the consequence of the fact that if the Ising model has long +range order, the free energy cost of the domain wall scales with its area. Moreover, in this phase +of the gauge theory the closed strings are small meaning that the string tension (the energy per +unit length of string) is finite. In the Ising model language, the string tension becomes surface +tension of the domain wall. +In the preceding subsection we discussed a Hamiltonian version of the duality. We will briefly +do the same in the case of the 2+1 dimensional Ising model. The hamiltonian of the Ising model +in a transverse field on a square lattice is +H2D−TFIM = − +� +r +σ1(r) − λ +� +r, j=1,2 +σ3(r)σ3(r + e j) +(108) +where r labels the sites of the square lattice and e j (with j = 1, 2) are the two (orthonormal) +primitive unit vectors of the lattice. Here too, at each site we have a two-level system (the spins), +σ1 and σ3 are Pauli matrices acting on these states at each site and λ is the coupling constant. +Just as in the 1D case, this system has two phases: an ordered phase for λ > λc and a disordered +phase for λ < λc, where λc is a critical coupling. This Hamiltonian is invariant under the global +Z2 symmetry generated by the global spin flip operator Q = � +r σ1(r) which commutes with the +Hamiltonian, [Q, H2D−TFIM] = 0. +Let is consider now a Z2 gauge theory on the dual of the square lattice. We will now define +a two-dimensional Hilbert space on each link, denoted by (˜r, j) (with j = 1, 2) of the dual lattice +with sites labelled by ˜r. We will further define at each link (˜r, j) the Pauli operators σ1 +j(˜r) and +σ3 +j(˜r). The Hamiltonian of the Z2 gauge theory is +HZ2gauge = − +� +˜r, j +σ1 +j(˜r) − g +� +˜r +σ3 +1(˜r)σ3 +2(˜r + ˜e1)σ3 +1(˜r + ˜e1 + ˜e2)σ3 +2(˜r) +(109) +where ˜e j = ǫjkek (with j, k = 1, 2). +Unlike the Hamiltonian of Eq.(108), the Hamiltonian of Eq.(109) has a local symmetry. Let +˜Q(˜r) be the operator that flips the Z2 gauge degrees of freedom on the four links that share the +site ˜r, +˜Q(˜r) = +� +j=1,2 +σ1 +j(˜r)σ1 +j(˜r − ˜e j) +(110) +29 + +These operators commute with each other, [ ˜Q(˜r), ˜Q(˜r′)] = 0, and commute with the Hamiltonian, +[HZ2gauge, ˜Q(˜r)] = 0. The Hilbert space of gauge-invariant states are eigenstates of ˜Q(r) with unit +eigenvalue, +˜Q(r)|Phys⟩ = |Phys⟩ +(111) +(for all r). This constraint is the the Z2 Gauss Law. The Hamiltonian HZ2gauge has two phases: a +confining phase for g < gc, and a deconfined phase for g > gc (where gc is a critical coupling). +The duality transformation is defined so that +σ1(r) = +� +plaquette(r) +σ3 +j(˜r) +(112) +is the product of the σ3 +j operators of the gauge theory on a dual plaquette centered at the site r, +and +σ1 +1(˜r) = σ3(r)σ3(r + e2), +σ1 +2(˜r) = σ3(r)σ3(r + e1) +(113) +These identities imply that the gauge theory constraint of the Z2 Gauss Law, Eq.(111), is satisfied. +This also means that duality is a mapping of the gauge-invariant sector of the Z2 gauge theory +to the Hilbert space of the Ising model in a transverse field. +Eq.(112) and Eq.(113) imply that the Hamiltonians H2D−TFIM and HZ2gauge are equal to each +other with the identification of the coupling constants g = 1/λ. Hence, the ordered phase of the +Ising model, λ > λc, maps onto the confining phase of the Z2 gauge theory and, conversely, the +disordered phase of the Ising model maps onto the deconfined phase of the gauge theory. +Eq.(113) also implies that the operator σ3(r) of the Ising model can be identified with the +operator +σ3(r) = +� +γ(r) +σ1 +j(˜r) +(114) +where the product is on the links of the dual lattice pierced by the path γ(r) on the direct lattice +ending at the site r. +While σ3(r) is of course just the order parameter of the Ising model, the dual operator defined +by Eq.(114) anti-commutes with the plaquette term of the Hamiltonian of Eq.(109) and creates an +Z2 flux excitation at the plaquette. With some abuse of language, this operator can be regarded as +creating a Z2 “magnetic monopole”. Since in the confining phase this operator has an expectation +value, we can regard this phase as a magnetic condensate. +We will now discuss the role of boundary conditions which is different in both theories. Let +us examine the behavior of the Z2 gauge theory with periodic boundary conditions. Periodic +boundary conditions means that the 2D space is topologically a 2-torus. It is straightforward to +show that the ground state in the confining phase is unique and insensitive to boundary condi- +tions. The physics of the deconfined phase is more subtle. We will now see that on a 2-torus +it has a four-fold degenerate ground state and that the degeneracy is not due to the spontaneous +breaking of a global symmetry. +Let γ1 and γ2 be two non-contractible loops on the torus along the directions 1 and 2 re- +spectively. Let us consider the (“electric”) Wilson loop operators along γ1 and γ2, W[γ1] = +� +(r, j)∈γ1 σ3 +j(r) and similarly for γ2. Similarly let us consider the (“magnetic”) ’t Hooft operators +˜W[˜γ1] and ˜W[γ2] defined on the non-contractible closed paths of the dual lattice ˜γ1 and ˜γ2, such +that ˜W[˜γ1] = � +˜γ1 σ1 +2(˜r) and ˜W[˜γ2] = � +˜γ2 σ1 +1(˜r). It is easy to see that the Wilson and’t Hooft +operators satisfy that W[γi]2 = ˜W[˜γ j]2 = I, and that +[W[γi], W[γ j] = [ ˜W[˜γi], ˜W[˜γ j]] = 0, +{W[γ1], ˜W[˜γ2]} = {W[γ2], ˜W[˜γ1]} = 0 +(115) +Let us consider the special limit of g → ∞. In this limit the Wilson loop operators W[γi] on +the two non-contractibleloops of the 2-torus commute with the Hamiltonian HZ2gauge of Eq.(109). +Therefore the eigenstates of the Hamiltonian can be chosen to be the eigenstates of these two Wil- +son loops. Since the Wilson loop operators are hermitian and obey W[γi]2 = I, the spectrum of +each loop is two-dimensional | ± 1⟩i (i = 1, 2) with eigenvalues ±1, respectively. At g → ∞ these +states are also eigenstates of HZ2gauge. Thus, the ground state of HZ2gauge is four dimensional. The +’t Hooft loops ˜W[˜γi] act as ladder operators in this restricted Hilbert space. +The conclusion is that, on a 2-torus, at least in the limit g → ∞, the ground state of HZ2gauge +is degenerate. However, this degeneracy is not due to the spontaneous breaking of a global +symmetry. Rather, it reflects the topological character of the theory. To see this, one can extend +this analysis to a theory to be on a more general two-dimensional and instead of a 2-torus we can +30 + +consider a surface with g handles (not to be confused with the coupling constant!), e.g. g = 0 for +a sphere (or disk), g = 1 for a 2-torus, g = 2 for a pretzel, etc. For each of these two-dimensional +surfaces the number different number of non-contractible loops is g, and the Wilson loops defined +on them commute with each other (and with HZ2gauge). Hence, in the limit g → ∞ HZ2gauge has +a ground state degeneracy of 2g which, clearly, depends only on the topology of the surface. We +conclude that, at least as g → ∞, the Hamiltonian HZ2gauge is in a topological phase. +However, is the exact degeneracy found in the limit g → ∞ a property of the entire deconfined +phase? For a finite system the expansion in powers of 1/g is convergent. On the other hand, in +the thermodynamic limit, L → ∞, the expansion has a finite radius of convergence with gc being +an upper bound. Let us consider the ground state at g = ∞ with eigenvalue +1 for the Wilson +loop W[γ1]. The degenerate state with eigenvalue −1 is created by the ’t Hooft loop ˜W[˜γ2], i.e. +| − 1⟩1 = ˜W[˜γ2] | + 1⟩1. The ’t Hooft operator is a product of σ1 operators on links along the +direction 1 crossed by the path ˜γ2 of the dual lattice, which involves (at least) L links. Then, it +takes L orders in the expansion in powers of 1/g to mix the state | + 1⟩1 with the state | − 1⟩1, +and this amplitude is of order 1/gL, which is exponentially small. The same argument applies to +the mixing between all four states. For a finite but large system of linear size L, the degeneracy +is lifted but the energy splitting is exponentially small in the system size. Hence, the topological +protection is a feature of the deconfined phase in the thermodynamic limit L → ∞. +We conclude that the deconfined phase of the Z2 lattice gauge theory is a topological phase. +This result extends to the case of a gauge theory with discrete gauge group Z, the cyclic group +of k elements. In general spacetime dimension D > 2 the Zk gauge theory also has a confined +and a deconfined phase and if k ≳ 4 it also has an intermediate “Coulomb phase”[125, 126]. In +section 9.4 we will see that the low-energy (IR) regime of the deconfined phase is described by +a topological quantum field theory known as the level k BF theory. +7. Bosonization +The behavior of one-dimensional electronic systems is of great interest in Condensed Matter +Physics for many reasons. One is that, even for infinitesimally weak interactions, these one- +dimensional metals violate the basic principles of the Landau Theory of the Fermi Liquid [13]. +A central assumption of Femi Liquid theory is that at low energies the excitation energy of the +fermion (electron) quasiparticle is always much larger than its width. Hence, at asymptotically +low energies the quasiparticle excitations become increasingly sharp. +A manifestation of this feature is that the quasiparticle propagator (the “Green function”) has +a pole on the real frequency axis with a finite residue Z. In one-dimensional metals this assump- +tion always fails since the residue Z vanishes, the Fermi field acquires a non-trivial anomalous +dimension, and the pole is replaced by a branch cut. To this date this is the best example of what +is called a ”non-Fermi liquid”. In one-dimensional metals this non-Fermi liquid is often called a +“Luttinger Liquid” [127]. A detailed analysis can be found in chapter 6 of Ref. [9], and in other +books. +Bosonization provides for a powerful tool to understand the physics of these non-trivial sys- +tems. Bosonization is a duality between a massless Dirac field in 1+1 dimensions and a (also +massless) relativistic Bose (scalar) field [128, 129, 130, 131]. In this context, bosonization is a set +of operator identities relating observables between two different (dual) continuum field theories. +These identities have a close resemblance to the Jordan-Wigner transformation which relates op- +erators of a theory of spinless (Dirac) fermions on a one-dimensional lattice to a theory of bosons +with hard cores (i.e. spins) on the same lattice [22]. +7.1. Dirac fermions in one space dimensions +To understand how these operator identities come about and what these fields mean in the +Condensed Matter context we will consider the simple problem of a system of non-interacting +spinless fermions c(n) on a one-dimensional chain of length L (with L → ∞), for simplicity with +periodic boundary conditions. The Hamiltonian is +H0 = −t +L +� +n=1 +c(n)†c(n + 1) + h.c. +(116) +In momentum space, and in the thermodynamic limit L → ∞, the Hamiltonian becomes +H0 = +� π +−π +dp +2π (ε0(p) − µ) c†(p)c(p) +(117) +31 + +where −π ≤ p ≤ π, µ is the chemical potential (which fixes the fermion number) and ε0(p) is the +(free) quasiparticle energy which, in this case, is +ε0(p) = −2t cos p +(118) +This simple system has a global internal symmetry +c′(n) = eiαc(n), +c′(n)† = e−iαc(c) +(119) +where α is a constant phase (with period 2π). This global symmetry reflects the conservation of +fermion number +NF = +� +n +c†(n)c(n) +(120) +The free fermion Hamiltonian of Eq.(116) is also invariant under lattice translations. +The ground state of this system is obtained by occupying all single particle states with energy +below the chemical potential, E ≤ µ ≡ EF, which defines the Fermi energy EF. In what follows +we will redefine the zero of the energy at the Fermi energy. In the thermodynamic limit the +one-particle states are labeled by momenta defined in the first Brillouin zone [−π, π). The ground +state |G⟩ of this system is obtained by filling up all single-particle states with momentum p in +the range [−pF, pF), where pF is the Fermi momentum. Hence, the occupied states have energy +ε0(p) ≤ EF. The ground state is +|G⟩ = +� +|p|≤pF +c†(p)|0⟩ +(121) +and is called the filled Fermi sea. We will further assume that the fermionic system is dense in +the sense that the occupied states are a finite fraction of the available states, i.e. we assume that +|EF| is a finite fraction of the bandwidth W = 4t. +The single-particle excitations have low energy if |ε0(p) − EF| ≪ |EF|. This range can only +be accessed by quasiparticles with momenta p close to ±pF. For states wit p close to pF we can +a linearized approximation and write +ε0(p) ≃ �F(p − pF) + . . . +(122) +and similarly for the states near −pF. Here �F = ∂ǫ0 +∂p +���pF = 2t sin pF is the Fermi velocity. Let q +being the momentum measured from pF (or −pF in the other case), This approximation is correct +in a range of momenta −Λ ≤ q ≤ Λ, where 2π +L ≪ Λ ≪ π. In this regime we can approximate +the lattice fields c(n) with two continuum right moving ψR(x) and left moving ψL(x) Fermi fields, +such that (with x = na0) +c(n) ∼ eipF xψR(x) + e−ipF xψL(x) +(123) +in terms of which the Hamiltonian takes the continuum form +Hcontinuum = +� +dx +� +ψ† +R(x)(−i�F)∂xψR(x) − ψ† +L(x)(−i�F)∂xψL(x) +� += +� dq +2πq�F +� +ψ† +R(q)ψR(q) + ψ† +R(q)ψR(q) +� +(124) +In what follows I will rescale time and space in such a way as to set �F = 1. +Eq.(124) is the Hamiltonian of a massless Dirac field in 1+1 dimensions. It describes the +effective low energy behavior of the excitations of the fermionic system defined on a lattice by +the Hamiltonian of Eq.(116). Here low energy means asymptotically close to the Fermi energy +(which we have set to zero) and momenta close to ±pF. We will see that this effective relativistic +field theory gives a complete description of the universal low energy physics encoded in the +microscopic model of Eq.(116) up to some subtle issues associated with what are called quantum +anomalies. +Let us denote by ψ(x) the bi-spinor field ψ(x) = (ψR(x), ψL(x)) and the 2 × 2 Dirac gamma- +matrices in terms of the three Pauli matrices are +γ0 = σ1, +γ1 = −iσ2, +γ5 = σ3 +(125) +which obey the Dirac algebra +{γµ, γν} = 2gµν +(126) +32 + +where gµν is the metric tensor in 1+1-dimensional Minkowski spacetime +gµν = +�1 +0 +0 +−1 +� +(127) +Using the standard notation ¯ψ(x) = ψ†(x)γ0 (where I left the spinor index α = 1, 2 implicit) +the Lagrangian density of the free massless Dirac fermion is +L = ¯ψi/∂ψ +(128) +where we have used the Feynman slash notation which denotes the contraction of a vector field, +say Aµ with the Dirac gamma matrices with a slash, Aµγµ ≡ /A. +Formally, the Dirac lagrangian of Eq.(128) (and the Dirac Hamiltonian of Eq.(124)) are in- +variant under two separate global transformations. It has a global U(1) (gauge) symmetry trans- +formation +ψ′(x) = eiθψ(x) +(129) +(where θ is constant) under which the two components of the bi-spinor ψ transform in the same +way +ψ′ +R(x) = eiθψR(x), +ψ′ +L(x) = eiθψL(x) +(130) +The massless Dirac theory is also invariant under a global gauge U(1) chiral transformation +ψ′(x) = eiθγ5ψ(x) +(131) +which in components it reads +ψ′ +R(x) = eiθψR(x), +ψ′ +L(x) = e−iθψL(x) +(132) +In general, if a system has a global continuous symmetry it should have a locally conserved +current and a globally conserved charge. This is the content of Noether’s Theorem. Since the +massless Dirac theory has these two global symmetries one would expect that it should have two +separately conserved currents. +The global U(1) symmetry has an associated current jµ = (j0, j1) given by +jµ = ¯ψγµψ +(133) +which is invariant under the global U(1) symmetry. In terms of the right and left moving Dirac +fields, ψR and ψL, the components of the current are +j0 = ψ† +RψR + ψ† +LψL, +j1 = ψ† +RψR − ψ† +LψL +(134) +This current is locally conserved and satisfies the continuity equation +∂µ jµ = 0 +(135) +It has an associated conserved global charge, which we will call fermion number +Q = +� ∞ +−∞ +dxj0(x) = +� ∞ +−∞ +dx +� +ψ† +RψR + ψ† +LψL +� +(136) +This is the continuum version of the conservation of fermion number NF of the free fermion +lattice model discussed above. +Since this current is conserved it can be coupled to an electromagnetic field through a term +in the Lagrangian +Lint = −ejµAµ +(137) +where Aµ is the electromagnetic vector potential, which in 1+1 dimensions has only two com- +ponents, Aµ = (A0, A1). Here e is a coupling constant which is interpreted as the electric charge. +This coupling amounts to making the U(1) symmetry a local gauge symmetry under which the +fields transform as follows +ψ′(x) = eiθ(x)ψ(x), +A′ +µ(x) = Aµ(x) + 1 +e∂µθ(x) +(138) +33 + +Now the conservation of fermion number becomes the conservation of the total electric charge +Qe = −eQ +(139) +In a continuum non-relativistic model the Fermi field is ψ(x) (ignoring spin), e.g. an electron +gas in one dimension such as a quantum wire, the analog of decomposition shown in Eq.(123) of +the electron field ψ(x) becomes +ψ(x) = eipF xψR(x) + e−ipF xψL(x) +(140) +The electron field ψ(x) transforms as ψ′(x) = eiαψ(x) under the global U(1) gauge symmetry of +Eq.(130). In particular, the local density operator ρ(x) = ψ†(x)ψ(x) is invariant under this sym- +metry. Under both decompositions of Eqs. (123) and (140), the local density operator becomes +ρ(x) = ψ† +R(x)ψR(x) + ψ† +L(x)ψL(x) + ei2pF xψ† +R(x)ψL(x) + e−i2pF xψ† +L(x)ψR(x) +(141) +Clearly the non-relativistic density operator ρ(x) is invariant under the global U(1) gauge sym- +metry. The same considerations applies to the lattice version of the density operator (the local +occupation number). +Thus, the density operator ρ(x) can be decomposed into a slowly varying part (the first two +terms in Eq.(141)) and the last two terms which oscillate with wave vectors Q = ±2pF. These +observations imply that we can express ρ(x) in the form +ρ(x) = ¯ρ + j0(x) + eiQxψQ(x) + e−iQxψ−Q(x) +(142) +This decomposition can be interpreted as a Fourier expansion of the density operator in term of +newly defined slowly-varying fields. In Eq.(142) Q = 2pF and ¯ρ is the average density, where we +assumed that the Dirac density operator j0(x) has vanishing expectation value (i.e. it is normal- +ordered). In Eq.(142) we defined the (bosonic) operators +ψQ(x) = ψ† +R(x)ψL(x), +ψ−Q(x) = ψ† +L(x)ψR(x) +(143) +which characterize the oscillatory component of the density. The operators ψ±Q(x) are the order +parameters of a charge density wave state in one dimension and Q is the ordering wavevector. +Repulsive interactions between the electrons can cause scattering process between the right +and left moving components to become relevant (in the Renormalization Group sense) leading +to the spontaneous breaking of translation invariance. The resulting state is known as a charge- +density-wave (CDW). In this state the operators ψ±Q(x) (or a linear combination of them) acquire +a non-vanishing expectation value, and the expectation value of the density operator ρ(x) has a +(static) modulated component, and the Dirac Hamiltonian density becomes +H = ψ† +R(x)(−i)∂xψR(x) − ψ† +L(x)(−i)∂xψL(x) + m +� +ψ† +R(x)ψL(x) + ψ† +L(x)ψR(x) +� +(144) +where m is the Dirac mass. The operator in the second term of Eq.(144) mixes the right and left +moving components of the Dirac spinor. This hermitian operator is known as the Dirac mass +term. In relativistic notation this operator is written +¯ψψ = ψ† +R(x)ψL(x) + ψ† +L(x)ψR(x) +(145) +When m � 0, i.e. when ⟨ ¯ψψ⟩ � 0, the fermionic spectrum has a mass gap, and the electronic +states with momenta ±pF have an energy gap. Alternatively we could have considered a state +with the in which the (hermitian) operator that has an expectation value is +i ¯ψγ5ψ = i +� +ψ† +R(x)ψL(x) − ψ† +L(x)ψR(x) +� +(146) +which is known as the γ5 mass term. In a more general CDW state both mass terms can be +present. +7.2. Chiral symmetry and chiral symmetry breaking +The CDW states we introduced have different symmetries. To see this let us observe that the +operators ψ±Q(x) transform non-trivially under a global U(1) chiral transformation +ψ′ +±Q(x) = e∓2iθψ±Q(x) +(147) +34 + +Upon substituting this transformation in the expansion of the density ρ(x) of Eq.(142), we see +that a chiral transformation is equivalent to a uniform displacement of the density operator by +θ/pF, +ρ(x) → ρ (x + θ/pF) +(148) +Under a chiral transformation with arbitrary angle θ, the Dirac and γ5 mass term operators trans- +form as an orthogonal transformation of a two-component vector. In particular for a chiral trans- +formation with θ = π/4, +¯ψψ �→ i ¯ψγ5ψ, +i ¯ψγ5ψ �→ − ¯ψψ +(149) +which is equivalent to a displacement of the density by 1/4 of the period of the CDW. In this +sense these two states are equivalent, as is the state with a more general linear combination. +The microscopic lattice model (and the continuum non-relativistic model) are invariant under +the spatial inversion symmetry x ↔ −x. This also implies a symmetry under the exchange of +right and left moving components of the Dirac spinor, ψR ↔ ψL. In the massless Dirac theory +this operation is equivalent to the multiplication of the spinor by the Pauli matrix σ1. In the +massless theory the multiplication by σ2 (followed by a chiral transformation with θ = π/2) has +the same effect. +These symmetries have an important effect on the fermionic spectrum. To understand what +they do let us consider the one-particle Dirac Hamiltonian with a Dirac mass m (jn momentum +space) +h = +�p +m +m +−p +� +(150) +This operator anti-commutes with the Pauli matrix σ2, {h, σ2} = 0. Let |E⟩ be an eigenstate of +the Hamiltonian h with energy eigenvalue E = +� +p2 + m2. Let us consider the state σ2|E⟩. It is +also an eigenstate of h but with energy −E, i.e. +hσ2|E⟩ = −σ2h|E⟩ = −E|E⟩, +⇒ σ2|E⟩ = | − E⟩ +(151) +Hence if |E⟩ is an eigenstate of energy E, then σ2|E⟩ is an eigenstate of energy −E. This means +that the spectrum is invariant under charge conjugation symmetry. Notice that under this opera- +tion the spinor transforms as +σ2 +�ψR +ψL +� += +�−iψL +iψR +� += e−iγ5π/2 +�ψL +ψR +� +(152) +In other words, this theory is invariant under charge conjugation C and parity P which, combined, +it implies that it is invariant under time-reversal T . +The same consideration applies in the case of a γ5 mass term in which case the one-particle +Dirac Hamiltonian now is +h = +� p +−im5 +im5 +−p +� +(153) +This Hamiltonian now anti-commutes with the Pauli matrix σ1 which also implies that the same +charge conjugation symmetry C, |E⟩ ↔ |−E⟩, is present in the spectrum of the case of a γ5 mass. +We can also repeat the argument on parity invariance P, which is now multiplication by σ1, Thus +the theory is invariant under time-reversal T . +However, if the theory has both a Dirac mass m and a γ5 mass m5 these symmetries are +broken. Indeed, the one-particle Dirac Hamiltonian now is +h = +� +p +m − im5 +m + im5 +−p +� += pσ3 + mσ1 + m5σ2 +(154) +which no longer has a spectral symmetry. In this case CP is broken and, hence, so it T since +CPT remains unbroken (as it should). +In a lattice system this is a symmetry transformation only if θ = πn/pF is a lattice displace- +ment, which restricts the allowed values of the chiral angle to be discrete. Although this is true +interactions play a significant role in the actual behavior. In fact, there are physical situations +in which an effective continuous symmetry actually emerges in this he infrared (long-distance) +limit. This is what happens when the CDW is incommensurate and, as we will see in the next +subsection, it slides under the action of an electric field. +In the case of a half-filled system (with only nearest-neighbor hopping matrix elements) the +Fermi wave vectors are ±π/2. In this case the allowed discrete chiral transformation has a chiral +35 + +angle π/2 under which ¯ψψ �→ − ¯ψψ (and similarly for i¯γ5ψ) corresponding by a translation by one +lattice spacing. The allowed four fermion operator is ( ¯ψψ)2. If the lattice fermions are spinless +this operator reduces to +( ¯ψψ)2 = −2 jR(x)jL(x) + lim +y→x ψ† +R(x)ψ† +L(x)ψL(y)ψR(y) + h.c. +(155) +where introduced the right and left moving (chiral) components of the current operator +jR(x) = 1 +2(j0(x) + j1(x)) = ψ† +R(x)ψR(x), +jL(x) = 1 +2(j0(x) − j1(x)) = ψ† +L(x)ψL(x) +(156) +The first term in Eq.(155) is known as the backscattering interaction and has scaling dimension +2. Hence, it is a marginal operator. The theory with only the first term is known in Condensed +Matter Physics as the Luttinger model and in High-Energy Physics at the (massless) Thirring +model. The second operator in Eq.(155) formally violates momentum conservation as its total +momentum is 4pF = 2π which is a reciprocal lattice vector and, as such, it is equivalent to +zero (mod 2π). Such an operator is not formally allowed in a (naive) continuum theory. This +operator is due to a lattice Umklapp process and breaks the formal continuous chiral symmetry +to a discrete Z2 subgroup. Although the naive scaling dimension of this operator is 2 (and hence +it is formally marginal). If the fermions are spinless, the leading operator actually vanishes and +the leading non-vanishing operator actually has dimension 4, which is irrelevant. However, as +shown above, backscattering processes of the form jR(x) jL(x) are part of this operator and are +exactly marginal. +If the interaction is strong enough the backscattering interaction can make the Umklapp op- +erator relevant. When this happens the fermionic system has a quantum phase transition to an +insulating state with a period 2 (commensurate) CDW state. On the other hand, for spin 1/2 +fermions the operator ( ¯ψψ)2 is allowed and is marginally relevant. If the interactions are re- +pulsive the resulting state is an antiferromagnetic N´eel state at quantum criticality, while for +attractive interactions it is a period 2 CDW. This is what happens in the 1D Hubbard model (see, +e.g. Ref. [9] for a detailed analysis). +There are two theories in relativistic systems which are closely related to this problem. One +if the Gross-Neveu model [132] which is a theory of N species of massless Dirac spinors with +Lagrangian density +LGN = ¯ψai/∂ψa + g( ¯ψaψa)2 +(157) +with a = 1, . . ., N (summation of repeated induces is implies). The spin-1/2 Hubbard model +corresponds to the case N = 2. This Lagrangian is invariant only under the discrete chiral +symmetry ¯ψaψa �→ − ¯ψaψa. This is a discrete, Z2, symmetry and as such it can be spontaneously +broken in 1+1 dimensions. For N ≥ 2, the resulting state has a (dynamically) broken Z2 chiral +symmetry and that there is a chiral condensate ⟨ ¯ψaψa⟩ � 0 corresponding to a period 2 CDW. +The other theory is known as the chiral Gross-Neveu model whose Lagrangian density is +LcGN = ¯ψai/∂ψa + g +� +( ¯ψaψa)2 − ( ¯ψaγ5ψa)2� +(158) +which has the full continuous chiral symmetry. For N = 1 this theory is equivalent to the Lut- +tinger model (and to the Gross-Neveu model if we ignore the umklapp term). For N ≥ 2 the chi- +ral symmetry is formally broken. If this were true this theory would violate the Mermin-Wagner +theorem. However a detailed study (most easily done using bosonization methods) shows that +instead of long range order the correlator of both mass terms decay as a power law as a function +of distance, consistent with the requirements by this theorem. +We close this subsection by noting that if the lattice model is not half filled but its density +is either incommensurate or has a higher degree of commensurability, say p/q, the chiral sym- +metry is actually continuous (in the incommensurate case) or effectively continuous since the +requisite umklapp terms are strongly irrelevant. However, if the lattice model is not at half filling +charge conjugation symmetry C is broken at the lattice scale (in the UV), where it is equivalent +to particle-hole symmetry, but it is recovered in the low-energy, IR, regime (up to irrelevant op- +erators). In this sense, both the continuous chiral symmetry and charge conjugation symmetry +can be regarded as emergent IR symmetries +7.3. The chiral anomaly +In section 7.2 we showed that the theory of massless Dirac fermions, in addition to a global +U(1) gauge gauge symmetry, has a second conservation law which we called a global U(1) chiral +36 + +symmetry, shown in Eq.(131). This symmetry implies that there is a locally associated chiral +current j5 +µ, given by +j5 +µ(x) = ¯ψ(x)γµγ5ψ(x) +(159) +which also satisfies a continuity equation +∂µ j5 +µ = 0 +(160) +and there is a globally conserved chiral charge Q5 +Q5 = +� ∞ +−∞ +dx j5 +0(x) = +� ∞ +−∞ +dx +� +ψ† +RψR − ψ† +LψL +� +(161) +It is easy to check that the Dirac current jµ and the chiral current j5 +µ are related by +j5 +µ = ǫµν jµ +(162) +where ǫµν is the second rank Levi-Civita tensor. +The simultaneous conservation of both currents jµ and j5 +µ in the massless Dirac theory implies +that the right and left moving densities jR and jL, defined in Eq.(156), should be separately +conserved. In fact, if the Dirac theory has a mass term +L = ¯ψi/∂ψ − m ¯ψψ +(163) +it is straightforward to show that +∂µ j5 +µ = 2mi ¯ψγ5ψ +(164) +which means that in the massive theory the axial current is not conserved. This is easy to under- +stand since the mass term mixes the right and left moving components of the Dirac spinor and, +hence, the right and left moving densities are not conserved. +What happens in the massless limit, m → 0, is more subtle. This problem was investigated +in the late 1960’s in 3+1 dimensions by S. Adler [133] and by J. S. Bell and R. Jackiw [134] +who were interested in the anomalous decay of a neutral pion into two photons, π0 → 2γ. +This process appears at third order of perturbation theory and it involves the computation of a +triangle diagram (a fermion loop). In 3+1 dimensions this process has a UV divergence which +needed to be regulated. These authors showed that it is not possible to find a regularization +in which both the Dirac (gauge) current jµ = ¯ψγµψ and the axial current j5 +µ = ¯ψγµγ5ψ are +conserved. In other words, if gauge invariance is preserved then the axial current is not and has +an anomaly and results in a non-conservation of the axial current, ∂µ j5 +µ � 0. On the other hand, +at least in the case of the physical gauge-invariant regularization, the obtained expression for +the anomaly in the axial current is universal, independent of the value of the UV regulator (the +cutoff). Sometime later G. ’t Hooft showed that in non-abelian gauge theories instant processes +also lead to anomalies and, furthermore, the result was also universal [96, 111]. Since the result +is universal and, hence, independent of the UV scale, this led to the concept of anomaly matching +conditions. +We will examine this problem in 1+1 dimensions (although it plays a key role in the theory of +topological insulators in three space dimensions [135]). As we noted above, the lattice model is +gauge-invariant and has only one conserved current. The conserved axial current appeared only +in the low-energy regime in which the lattice model is described by a theory of massless Dirac +fermions. To understand this problem we will consider the theory of fermions in 1+1 dimen- +sions coupled to a U(1) gauge field field. In the presence of a background (i.e. not quantized) +electromagnetic field in the A0 = 0 gauge the free fermion Hamiltonian of Eq.(116) becomes +H[A] = −t +L +� +n=1 +c†(n)eieA(n,t)/ℏc(n) + h.c. +(165) +An uniform and constant electric field is E is represented by a vector potential A = −cEt (with +c being the speed of light). The net effect of this gauge field is to shift of the momentum of the +fermion quasiparticles p → p + ecEt/ℏ or, what is the same, to displace the fermion dispersion +relation in momentum by the ecEt/ℏ. This means that the Fermi points are also shifted by that +amount, pF → pF + ecEt/ℏ and −pF → −pF + ecEt/ℏ. This means that the single particle states +between pF and pF +ecEt/ℏ that were empty for E = 0 are now occupied and the states between +37 + +−pF and −pF + ecEt/ℏ that were occupied for E = 0 are now empty. This means that number +of right-moving fermions is increasing at a rate of ecE/ℏ and that the number of left-moving +fermions decreases at the same rate. This results in a net current. Throughout this process the +total number of fermions is not changed, gauge invariance is satisfied, but the number of right +and left moving fermions are not separately concerned. Notice that this is an effect that involved +the entire Femi sea but the net effect is at low energies. +Let us see now how this plays out in the effective Dirac theory. The massless Dirac La- +grangian density in the background of an unquantized electromagnetic field Aµ is +L = ¯ψ(i/∂ − e /A)ψ +(166) +Since there is no mass term the Dirac equation still decouples into two equations, for the right +and left moving components of the Dirac spinor. In the A0 = 0 gauge they are +i∂0ψR = (−i∂1 − A1)ψR, +i∂0ψL = (i∂1 − A1)ψL +(167) +In the temporal gauge, A0 = 0, a uniform electric field E = ∂0A1, and A1 increases linearly with +time. As A1 increases, the Fermi momentum pF (which is equal to the Fermi energy EF) also +increases at the rate eE. The density of states of a system of length L is L/(2π). So, the rate of +change of the number of right-moving fermions is +dNR +dx0 += e +2πE +(168) +where we defined NR = +� L +0 dx jR and similarly for NL. If the UV regulator of the theory is +compatible with gauge invariance, then the total fermion number must be conserved and the total +vacuum charge must remain equal to zero, +Q = +� L +0 +dx j0(x) = NR + NL = 0 +(169) +Thus, if NR increases, then NL must decrease by the same amount. Or, equivalently, the electric +field E creates an equal number of particles NR and of antiparticles ¯NL = −NL. +On the other hand, the chiral charge Q5 = NR − NL must increase at the rate +dQ5 +dx0 += dNR +dx0 ++ d ¯NL +dx0 += e +π E +(170) +Again, the details of the UV regularization do not matter, only that it is gauge-invariant. We can +also interpret Eq.(170) as the rate of particle-antiparticle pair creation by an electric field. +In a relativistic notation these results are expressed as +∂µ j5 +µ = e +2πǫµνFµν +(171) +Hence, the formally conserved current j5 +µ has an anomaly and is not conserved due to quantum +effects. Since it is not conserved, we cannot gauge the chiral symmetry. In the next subsection +we will see that the the anomaly is closely related with bosonization. +7.4. Bosonization, anomalies and duality +We will reexamine the problem at hand from the point of view of the fermionic currents of +the Dirac theory jµ as operators. Since the currents obey the continuity equation, ∂µ jµ = 0 one +expects that this would imply that it may be possible to write them as a curl of a scalar field, i.e. +jµ(x) = ǫµν∂νφ(x) where φ(x) should be a scalar field. Since this should be an operator identity +we will need to understand how the currents act on the physical Hilbert space. +The Fermi-Bose mapping in 1D systems is closely related to the chiral anomaly we just +discussed. Bosonization of a system of 1+1 dimensional massless Dirac fermions is a set of +operator identities understood as matrix elements of the observables in the physical Hilbert space. +These identities were first derived by Daniel Mattis and Elliott Lieb [128], based on earlier work +by Julian Schwinger [136]. These identities were rediscovered (and their scope greatly expanded) +by Alan Luther and Victor Emery [129], by Sidney Coleman [131], by Stanley Mandelstam +[130], and by Edward Witten [137]. A non-abelian version of bosonization was subsequently +derived by Witten [119]. +38 + +The physical Hilbert space is defined as follows. Let |FS⟩ ≡ |0⟩ denote the filled Fermi +sea. In what follows we will assume that the physical system is macroscopically large abd that +local operators create the physical states by acting finitely on the filled Fermi sea. Physical +observables, such as the right and left moving densities jR(x) and jL(x), need to be normal- +ordered with respect to the physical vacuum state, the filled Fermi (Dirac) sea |0⟩. The normal +ordered densities are : jR(x) : ≡ jR(x) − ⟨0|jR(x)|0⟩ and : jL(x) : ≡ jL(x) − ⟨0|jR(x)|0⟩. Since +the densities are products of fermion operators they need to be defined as a limit in which the +operators are separated by a short distance η. Crucial to this construction is that the computation +the expectation values be regularized in such a way that the charge (gauge) current jµ(x) is locally +conserved and satisfies the continuity equation ∂µ jµ = 0 as an operator identity. In what follows +all expectation values will refer to the filled Fermi sea state |0⟩. +The propagators of the right and left moving Fermi fields are given by +⟨ψ† +R(x0, x1)ψR(0, 0)⟩ = +−i +2π(x0 − x1 + iǫ), +⟨ψ† +L(x0, x1)ψL(0, 0)⟩ = +i +2π(x0 + x1 + iǫ) +(172) +The expectation value of the currents at a space location x1 are +⟨jR(x1)⟩ = lim +η→0⟨ψ† +R(x1 + η)ψR(x1 − η)⟩ = +i +4πη, +⟨jL(x1)⟩ = lim +η→0⟨ψ† +L(x1 + η)ψL(x1 − η)⟩ = −i +4πη +(173) +which are divergent at short distances. It follows that the normal-ordered right and left moving +current densities satisfy the equal-time commutation relations +[jR(x1), jR(x′ +1)] = − i +2π∂1δ(x1 − x′ +1), +[jL(x1), jL(x′ +1)] = + i +2π∂1δ(x1 − x′ +1) +(174) +These identities imply that the normal-ordered space-time components of the current jµ = (j0, j1) +satisfy the equal-time commutation relations +[j0(x1), j1(x′ +1)] = − i +π∂1δ(x1 − x′ +1), +[j0(x1), j0(x′ +1)] = [ji(x1), j1(x′ +1)] = 0 +(175) +The non-vanishing right-hand sides of these commutators are known as Schwinger terms. These +identities define the the U(1) (Kac-Moody) current algebra. +We should note that in a theory of non-relativistic Fermi fields ψ(x, t) in all dimensions, the +charge density ρ(x) and the current operators j(x) = +1 +2i +� +ψ†(x)▽ψ(x) − ▽ψ†(x)▽ψ(x) +� +satisfy a +similar expression (also at equal times) +[ρ(x), jk(x′)] = −i e2 +mc2 ∂k +�δ(x − x′)ρ(x)� , +[ρ(x), ρ(x′)] = 0, +[jk(x), jl(x′)] = 0 +(176) +In one dimension, and in the regime in which the fermions have a macroscopic density so that +ρ(x) ≃ ⟨ρ(x)⟩, after a multiplicative rescaling of the operators, the non-relativistic identities of +Eq.(176) are equivalent to the U(1) current algebra of Eq.(175). +The U(1) current algebra of Eq,(175) is reminiscent of the equal-time canonical commutation +relations of a scalar field. Indeed, if φ(x) is a scalar field and Π(x) = ∂0φ(x) is its canonically +conjugate momentum, then they obey the equal-time canonical commutation relations +[φ(x1), Π(x′ +1)] = iδ(x1 − x′ +1) +(177) +We can then identify the charge density j0 and the current density j1 with the scalar field operators +j0(x) = +1√π∂1φ(x), +j1(x) = − 1√πΠ(x) = − 1√π∂0φ(x) +(178) +which obey the U(1) current algebra of Eq.(175) as a consequence of the canonical commutation +relations, Eq. (177). Furthermore, we can rewrite Eq.(177) in the Lorentz covariant form +jµ(x) = +1√πǫµν∂νφ(x) +(179) +which is clearly consistent with the local conservation of the current jµ, +∂µ jµ(x) = 0 +(180) +39 + +Let us examine now the question of the conservation of the chiral current j5 +µ. In Eq.(162) +we showed that the gauge current and the chiral current are related by j5 +µ = ǫµν jν. Therefore the +divergence of the chiral current is +∂µ j5 +µ = ǫµν∂µ jν = +1√πǫµνǫνλ∂µ∂λφ = − 1√π∂2φ +(181) +where we used the identification of the gauge current in terms of the scalar field φ, Eq.(179). +Therefore we conclude that +∂µ j5 +µ = 0 ⇔ ∂2φ = 0 +(182) +This equation states that the chiral current as an operator identity is conserved if and only if the +field φ is a free massless scalar field, whose Lagrangian density is +LB = 1 +2(∂µφ)2 +(183) +In Eq.(166) we considered the free massless Dirac Lagrangian coupled to a background (not +quantized) gauge field Aµ through the usual minimal coupling which here is Lint = −ejµAµ. using +the bosonization identity for the gauge current, Eq.(179) we see that the Lagrangian density of +the bosonized theory now becomes +LB[A] =1 +2(∂µφ)2 − +e√πǫµν∂νφ(x) Aµ(x) +≡1 +2(∂µφ)2 + J(x)φ(x) +(184) +where the source J(x) is +J(x) = +e√πǫµν∂νAµ(x) = +e +√ +4π +F∗(x) +(185) +where F∗(x) = ǫµνFνµ(x) is the (Hodge) dual of the field strength Fµν. Hence, F∗ is (essentially) +the source for the scalar field φ(x). This implies that the equation of motion of the scalar field +must be +− ∂2φ(x) = J(x) = +e√πǫµν∂νAµ +(186) +Retracing our steps we find that the chiral current j5 +µ obeys +∂µ j5 +µ = − 1√π∂2φ = e +2πǫµνFµν +(187) +which reproduces the the chiral anomaly given in Eq.(170). +These results suggest that the theory of a free massless Dirac spinor must be equivalent to +the theory of the free massless scalar field. This statement is known as bosonization. However +for this identification to be correct there must be an identification of the Hilbert spaces and of all +the operators of each theory. We will not do this detailed analysis here but we will highlight the +most significant statements. +Let us begin with the fermion number of the Dirac theory. Consider a system of fermions +of total length L. Using the bosonization identity of Eq.(179) we find that the fermion number +NF ≡ Q is given by +NF = +� L +0 +dx1 j0(x0, x1) = +1√π +� L +0 +dx1 ∂1φ(x0, x1) = +1√π(φ(x0, L) − φ(x0, 0)) +(188) +Thus, the vacuum sector of the Dirac theory, with NF = 0, corresponds to the theory of the scalar +field with periodic boundary conditions, φ(x1 = 0) = φ(x1 = L). Furthermore, since the fermion +number is quantized, NF ∈ Z, changing the fermion number is the same as twisting the boundary +conditions of the scalar so that +φ(x1 + L) = φ(x1) + √πNF +(189) +In String Theory [104] the scalar field is interpreted as the coordinate of a string. Compactifying +the space where the string lives to be a circle of radius R means that the string coordinate is +defined modulo 2πRn, where n is an integer. We see that the condition imposed by Eq.(189) +40 + +is equivalent to say that the scalar field is compactified and that the compactification radius is +R = 1/ +√ +4π. This identification also imposes the restriction that the allowed operators of the +scalar field must obey the identification +φ(x) ∼ φ(x) + 2πRn +(190) +as an equivalency condition. +The simplest bosonic operators that obey the compactification condition are the vertex oper- +ators Vα(x), +Vα(x) = exp(iαφ(x)) +(191) +The compactification condition then requires that the allowed vertex operators should have α = +n/R = +√ +4π n, where n is an integer. Since the propagator of the scalar field in 1+1-dimensional +(Euclidean) spacetime is +G(x − x′) = − 1 +2π ln +�|x − x′| +a +� +(192) +where a is a short-distance cutoff, we find that the scaling dimension of the vertex operator is +∆α = α2/(4π) = n2. We will see shortly that the vertex operator with α = +√ +4π is essentially the +Dirac mass operator (which has scaling dimension 1). +The free massless scalar field can be decomposed into right and left moving components, φR +and φL respectively, +φ = φR + φL, +ϑ = −φR + φL +(193) +where +ϑ(x0, x1) = +� x1 +−∞ +dx′ +1Π(x0, x′ +1) +(194) +is the Cauchy-Riemann dual of the field φ(x) since they satisfy the Cauchy-Riemann equation +∂µφ = ǫµν∂νϑ +(195) +The right and left moving component of the Dirac spinor are found to have the bosonized expres- +sion [130] +ψR(x) = +1 +√ +2πa +: exp(i +√ +4πφR(x)) :, +ψL(x) = +1 +√ +2πa +: exp(−i +√ +4πφL(x)) : +(196) +It is easy to check that the propagators of these operators agree with the expressions given in +Eq.(172), and that they have scaling dimension 1/2 and spin 1/2. +How does a chiral transformation act on the scalar field? A chiral transformation by an angle +θ, c.f. Eq.(132), acts on the right and moving fermions as ψ′ +R = exp(iθ)ψR and ψ′ +L = exp(−iθ)ψL. +From Eq.(196) we see that the right and left moving components of the scalar field transform as +φ′ +R = φR + θ/ +√ +4π and φ′ +L = φL + θ/ +√ +4π. This means that a chiral transformation by an angle +θ of the Dirac fermion by an angle θ is equivalent to a translation (a shift) of the scalar field +φ′ = φ + 2θ/ +√ +4π. +We can use the Operator Product Expansion discussed in section 3.3.2 to show that the +fermion mass terms ¯ψψ and i ¯ψγ5ψ are given by the following identifications +¯ψψ = +1 +2πa : cos( +√ +4πφ) :, +i ¯ψγ5ψ =: sin( +√ +4πφ) : +(197) +These operators have scaling dimension 1 and transform properly under chiral transformations. +These identifications imply that a theory of free massive Dirac fermions +LD = ¯ψi/∂ ψ − m ¯ψψ +(198) +is equivalent to the sine-Gordon field theory [131] whose Lagrangian is +LSG = 1 +2(∂µφ)2 − g : cos( +√ +4πφ) : +(199) +where g = m/(2πa). +Given the central role played by the current algebra identities of Eqs. (175) one may wonder +if a similar approach might apply in higher dimensions. Schwinger terms in current algebra play +an important role in relativistic field theories. However in higher dimensions their structure is +41 + +more complex and does not lead to identities of the type we have discussed. The reason at the +root of this problem is largely kinematical. The Bose (scalar) field φ is qualitatively a bound +state, a collective mode in the language of Condensed Matter Physics. In 1+1 dimensions this +collective mode exhausts the spectrum at low energies due to the strong kinematical restriction +on one spatial dimension. +The equivalency between the theory of free massive Dirac fermions and the sine-Gordon +theory is an example of the power of bosonization. On the Dirac side the mapping the theory is +free and its spectrum is well understood. But on the sine-Gordon side the theory is non-linear. +In fact in the sine-Gordon theory the fermions are essentially solitons, domain walls of the scalar +field. For these and many other reasons that we do not have space here bosonization plays a +huge role in understanding the non-perturbative behavior of systems both in Condensed Matter +Physics and in Quantum Field Theory in 1+1 dimensions. We will see in section 11.3.1 that to +an extent some of these ideas can and have been extended to relativistic systems and classical +statistical mechanical systems in 2+1 dimensions. +8. Fractional Charge +8.1. Solitons in one dimensions +We begin by returning to the equivalency between the theory of free massive Dirac fermions +and sine-Gordon theory, in 1+1 dimensions. In Eq.(188) we showed that the boundary conditions +of the compactified scalar field φ(x) are determined by the fermion number NF of the dual Dirac +theory, and that the vacuum sector of the Dirac theory maps onto the sine-Gordon theory with +periodic boundary conditions. We will now examine the sector with one fermion, NF = 1. This +sector of the Dirac theory maps onto the sine-Gordon theory with twisted boundary conditions, +φ(L) − φ(0) = √π +(200) +The Hamiltonian of the sine-Gordon theory is +HSG = +� ∞ +−∞ +dx +�1 +2Π2(x) + 1 +2 (∂xφ(x)) + g cos +� √ +4πφ(x) +�� +(201) +In the sector with periodic boundary conditions the classical ground states are static and uniform +configurations that minimize the potential energy. Since the potential energy is a periodic func- +tional of φ(x) the classical minima are at φn(x) = (n + 1/2) √π, where n is an arbitrary integer. +The classical energy of these ground states is extensive and is given by Egnd = −gL where L is +the linear size of the system. +The classical ground state in the twisted sector is a domain wall (or soliton) which interpo- +lates between the static and uniform ground states φ(x) = ± √π/2. The classical ground state in +this sector is the static solution of the Euler-Lagrange equation +d2φ +dx2 = −2g √π sin +� +2 √πφ(x) +� +(202) +such that asymptotically satisfies limx→±∞ = ± √π/2. The solution is the classical soliton con- +figuration +φ(x) = +2√π tan−1 � +exp(2 √πg(x − x0)) +� +− +√π +2 +(203) +The soliton solution represents a domain wall between two symmetry-related classical ground +states with φ = ± √π/2. +The energy of the soliton (measured from the energy of the ground state in the trivial sector) +is finite and is given by +Esoliton = 4 +� +g +π +(204) +where x0 is a zero mode of the soliton solution and represents its coordinate. By coupling the +bosonized theory to a weak electromagnetic field Aµ, as given in Eq.(184), it is easy to check that +it has electric charge −e and, in this sense represents the electron. Then the identities of Eq.(196) +can be used that as a quantum state it is indeed a fermion. +42 + +8.2. Polyacetylene +In section 7.4 we saw that solitons of a scalar field can be regarded as being equivalent to +electrons, fermions with charge −e. We will now see that in a theory of fermions coupled to +a domain wall of a scalar field, the soliton carries fractional charge. This problem has been +extensively studied in one-dimensional conductors such as polyacetylene, in particular by the +work of Wu-Pei Su, J. Robert Schrieffer and Alan Heeger [138] and by Roman Jackiw and J. +Robert Schrieffer [139]. In quantum field theory this problem was first discussed by Roman +Jackiw and Claudio Rebbi [140] and by Jeffrey Goldstone and Frank Wilczek [141]. +In section 7.1 we showed that the physics of lattice fermions in one dimension at low energies +is well described by a theory of massless Dirac fermions. In a one-dimensional conductor, such as +polyacetylene, the fermions couple to the lattice vibrations (phonons). Su, Schrieffer and Heeger +(SSH) [138] proposed a simple model in which the electrons couple to the lattice vibrations +through a modulation of the hopping amplitude between two consecutive sites n and n + 1, +instead of being a constant t, becomes tn,n+1 = t − g(un+1 − un), where un is the displacement of +the ion (a CH group in polyacetylene) at site n from its classical equilibrium position and g is +the electron-phonon coupling constant. In polyacetylene there number of electrons (which are +spin-1/2 fermions) is equal to the number of sites of the lettuce and the electronic band is half- +filled. At half filling this simple band structure is invariant under a particle-hole transformation. +If the coupling to the lattice vibrations is included this symmetry remains respected provided the +displacements change sign un → −un for all lattice sites. In polyacetylene the lattice dimerizes +(a process known as a Peierls distortion) and the discrete translation symmetry by one lattice +spacing is spontaneously broken: the system becomes a period 2 CDW on the bonds of the +lattice. The broken symmetry state is still invariant under a particle-hole transformation. +The effective field theory of this system is a theory of two Dirac spinors ψα,σ(x), where +α = 1, 2 denotes right and left-moving fermions, and σ =↑, ↓ are the two spin polarizations. The +Lagrangian density of this system is +L = ¯ψσi/∂ψσ − gφ(x) ¯ψσ(x)ψσ(x) − 1 +2φ(x)2 +(205) +The real scalar field φ(x) represents the distortion field of the polyacetylene chain. Here we will +assume that the chains has spontaneously distorted and we will regard the scalar field as static +and classical. This is a good approximation since the masses of the CH complexes is much +bigger than the electron mass. This continuum model is due to Takayama, Lin-Liu and Maki +[142] and further developed by Campbell and Bishop [143, 144]. Many of the results fund in +this (adiabatic) approximation remain qualitatively correct upon taking into account the quantum +dynamics of the chain, even in the limit in which the ions are treated as being “light” (provided +the spin of the fermions is taken into account) [145, 146]. +In the field theory the discrete symmetry of displacements by one lattice spacing becomes +the Z2 symmetry φ → −φ. This is a symmetry of the electron phonon system once combined +with the discrete chiral transformation ψ → γ5ψ under which ¯ψψ → − ¯ψψ. The ground state is +two fold degenerate ±φ0 with +φ0 = 2Λ�F +g +exp +� +−π�F +g2 +� +(206) +and the Dirac fermion (the electron) has a exponentially small mass, m = gφ0. +8.3. Fractionally charged solitons +Jackiw and Rebbi showed that the 1+1-dimensional classical φ4 theory has the following +soliton solution which interpolates between the two classically ordered states at ±φ0 [140] +φ(x) = φ0 tanh +� x − x0 +ξ +� +(207) +where ξ is the correlation length of φ4 theory, and x0 is the (arbitrary) location of the soliton. +They further showed that, when coupled to a theory of massless relativistic fermions through a +Yukawa coupling, as in Eq.(205), this soliton carries fractional charge. The argument goes as +follows. The one-particle Dirac Hamiltonian for a Dirac fermion with a position-dependent mass +m(x) is +H = −iσ1∂x + m(x)σ3 = +�m(x) +−i∂x +−i∂x +−m(x) +� +(208) +43 + +where m(x) = gφ(x), with φ(x) being the soliton solution of Eq.(207). This Hamiltonian is her- +mitian and real. Furthermore, this Hamiltonian anti-commutes with the Pauli matrix σ2. This +implies that for every positive-energy state |E⟩ with energy +E there is a negative-energy eigen- +state with energy −E given by σ2|E⟩. Hence, the spectrum is particle-hole symmetric (or, what +is the same, charge-conjugation invariant). In addition, and consistent with charge-conjugation +symmetry, the Hamiltonian of Eq.(208) has state with E = 0, a zero-mode, with a normalizable +spinor wave function +ψ0(x) = +1√ +2 +�−i +1 +� +exp +� +−sgn(m) +� x +0 +dx′ m(x′) +� +(209) +which exists for an arbitrary function m(x) which changes sign once at some location (which +we took to be x0 = 0). Jackiw and Rebbi further showed that the soliton (anti-soliton) carries +fractional charge +Q = ∓e +2 +(210) +This result follows from the spectral asymmetry identity of the density of states ρS (E) in the +presence of the soliton +Q = −e +2 +� ∞ +0 +(ρS (E) − ρS (−E)) = −e +2η +(211) +where η is the spectral asymmetry of the Dirac operator in the soliton background, and it is known +as the APS η-invariant of Atiyah-Patodi-Singer [147]. Given the one-to-one correspondence +that exists between positive and negative energy states in the spectrum, the spectral asymmetry +follows from the condition that the zero mode be half-filled, which is required by normal-ordering +or, what is the same, by charge neutrality. Another way to understand this result is that adding (or +removing) a fermion of charge −e results in the creation of a soliton-antisoliton pair, with each +topological excitation carrying half of the charge of the electron. In other words, in the dimerized +phase the electron is fractionalized. In this analysis we ignored the spin of the electron. If we +take it into account the spin degree of freedom the soliton is instead a boson with charge ∓e. +There is an alternative, complementary, way to think about the charge of the soliton. Gold- +stone and Wilczek [141] considered a theory in which the (massless) Dirac fermion is coupled to +two real scalar fields, ϕ1 and ϕ2, with Lagrangian +L = ¯ψi/∂ψ − gϕ1 ¯ψψ − igϕ2 ¯ψγ5ψ ≡ ¯ψi/∂ψ − g|ϕ| ¯ψ exp(iθγ5)ψ +(212) +where |ϕ|2 = ϕ2 +1 + ϕ2 +2 and θ = tan−1(ϕ2/ϕ1). They considered a soliton in which gϕ1 = m is the +constant (in space) Dirac mass and ϕ2 winds slowly between two values ±ϕ0 for x → ±∞. In +this theory the one-particle Dirac Hamiltonian is +H = −iσ1∂x + gϕ1σ3 + gϕ2σ2 +(213) +which is hermitian and complex and, hence, it violates CP invariance. +A perturbative calculation of the induced (gauge-invariant) current jµ, which is given by the +triangle diagram of a fermion loop with two gauge field insertions and a coupling of the scalar +fields, yields the result (with a = 1, 2) +⟨jµ(x)⟩ = 1 +2πǫµνǫab +ϕa∂νϕb +|ϕ|2 += 1 +2πǫµν∂νθ +(214) +which is locally conserved. Notice, however, that the induced axial current j5 +µ is not conserved, +∂µ⟨j5 +µ⟩ = − 1 +2π∂2θ � 0 and, hence, this current is anomalous. We can now compute the total +charge accumulated as the soliton is created adiabatically to be given by the Goldstone-Wilczek +formula +Q = −e∆θ +2π +(215) +where ∆θ = θ(+∞) − θ(−∞). Since limx→±∞ ϕ2(x) = ±ϕ0, we obtain the result +Q = − e +π tan−1 �gϕ0 +m +� +(216) +In the limit m = gϕ → 0, where CP (or T) invariance is recovered, we get +lim +m→0 Q = −e +2 +(217) +44 + +which is the Jackiw-Rebbi result for the fractional charge of a soliton of Eq.(210). +The results for the fractional charge of the soliton can also be derived using the bosonization +identities of section 7.4. Indeed, the bosonized expression for the Lagrangian of Eq.(212) is +L = 1 +2 +� +∂µφ +�2 − g|ϕ| +2πa cos +� √ +4πφ − θ +� +(218) +Deep in the phase in which the Bose field φ is massive, the non-linear term in Eq.(218) locks this +field to the chiral angle θ, i.e. +φ = +1 +√ +4π +θ +(219) +However, the bosonization identities also tell us that the gauge current jµ is given by the curl of +the scalar field φ. Therefore, in this state the current jµ is +jµ = +1√πǫµν∂νφ = 1 +2πǫµν∂νθ +(220) +which is the same as the Goldstone-Wilczek result of Eq.(214). That these two seemingly differ- +ent approaches yield the same result is not accidental as they both follow from the axial anomaly. +9. Fractional Statistics +A fundamental axiom of Quantum Mechanics is that identical particles are indistinguish- +able [148, 149]. In non-relativistic Quantum Mechanics this leads to the requirement that the +quantum states of a system of identical particles must be eigenstates of the pairwise particle ex- +change operator. Since two exchanges are equivalent to the identity operation this implies that the +states must be even or odd under pairwise exchanges. This result, in turn, implies that particles +can be classified as either being bosons (whose states are invariant under pairwise exchanges) +or fermions (whose states change sign under pairwise particle exchanges). A consequence is +that bosons obey the the Bose-Einstein (and can condense into a single particle state) whereas +fermions obey the Fermi-Dirac distribution and must obey the Pauli exclusion principle. It is +an implicit assumption of this line of reasoning that all relevant states of a system of identical +particles can be efficiently represented by a (suitably symmetrized or antisymmetrized) product +state. +This classification is present at an even deeper level in (relativistic) Quantum Field Theory +where locality, unitarity and Lorentz invariance require that the fields be classified as represen- +tations of the Lorentz group and obey the Spin-Statistics Theorem [150]. The Spin-Statistics +Theorem is actually an axiom of local relativistic Quantum Field theory which requires that +fields that transform with an integer spin representation of the Poincar´e group (i.e. scalars, gauge +fields, gravitational fields, etc) must be bosons while fields that transform with a half-integer +spin representation (i.e. Dirac spinors, etc) must be fermions. This spin-statistics connection is +intrinsic to the construction of String Theory [104]. +Given these considerations there was a general consensus that fermions and bosons were +the only possible types of statistics. Nevertheless several exceptions to this rule were known +to exist. One is the construction of the magnetic monopole in 3+1 dimensional gauge theory +by Tai-Tsun Wu and Chen-Ning Yang who showed that a scalar coupled to a Dirac magnetic +monopole behaves as a Dirac spinor [151]. This was an early example of statistical transmutation +by coupling a matter field to a non-trivial configuration of a gauge field. As we will note below, +the construction of anyons (particles with fractional statistics) in 2+1 dimensions has a close +parentage to the Wu-Yang example. Examples of statistical transmutation were known to exist +in 1+1 dimensional theories where a system of hard-core bosons was shown to be equivalent to a +theory of free fermions using the Jordan-Wigner transformation [22] which represents a fermion +as a composite operator of a hard-core boson and an operator that creates a kink (or soliton). +This construction also underlies the fermion-boson mapping in 1+1 dimensional field theories +[128, 129, 131, 130] that we discussed in section 7.4. Finally, in the late 1970s it was found that +1+1 dimensional ZN spin systems harbor operators known as parafermions which obey the same +algebra shortly afterwards found to be obeyed by anyons in 2+1 dimensions [152]. +45 + +9.1. Basics of fractional statistics +Jon Magne Leinaas and Jan Myrheim wrote an insightful paper in 1977 in which they exam- +ined the structure of the configuration space of the histories of a system of N identical particles +[153]. Using the Feynamn path-integral approach, they showed that if the worldlines of the iden- +tical particles are not allowed to cross, then the configuration space is topologically non-trivial. +Through a detailed analysis they showed that the three and higher dimensions under a pairwise +particle exchange the states must be either even or odd and hence the particles are either bosons +or fermions. Leinaas and Myrheim also showed that in one and two space dimensions the wave +functions can change by a phase, nowadays known as the statistical angle. In retrospect this re- +sult could have been anticipated (but was not) in an earlier paper by Michael G. G. Laidlaw and +C´ecile Morette DeWitt [154] who did a similar analysis of the configuration space of identical +particles in the Feynman path integral. +Frank Wilczek generalized the Aharonov-Bohm effect [155] to describe the quantum me- +chanics of composite objects made of electric charge and magnetic flux in two space dimensions +[156]. Wilczek showed that composite objects made of a non-relativistic particle of charge q +bound to a magnetic flux of (magnetic) charge Φ behaves as an object with fractional angu- +lar momentum qΦ/2π. Here Φ is measured in units of the flux quantum 2π (in units in which +ℏ = c = e = 1). Furthermore, in a subsequent paper Wilczek [157] showed that, upon an +adiabatic process in which the two composites exchange positions without their worldlines co- +inciding, the wave function of two identical flux-charge composites changes by a phase factor +exp(±iqΦ). For instance if Φ = π (i.e. a half-flux quantum) the acquired phase is equal to +exp(iπ) = −1. thus if the particle was a boson, the composite becomes a fermion and viceversa. +Wilczek coined the term anyon to describe the behavior of an arbitrary charge-flux composite. +Furthermore, this construction also implies that not only fractional statistics but also fractional +spin, consistent with a generalization of the Spin-Statistics Theorem. In other terms, “flux- +attachment” implies fractional statistics (and fractional spin). Clearly Wilczek’s construction +gave an explicit physical grounding to the general arguments of the 1977 paper by Leinaas and +Myrheim. It is worth note the close analogy between this construction in 2+1 dimensions and the +Wu-Yang construction in 3+1 dimensions (whose consistency with the Spin-Statistics Theorem +had been shown earlier on by Alfred Goldhaber [158]). +In this description the statistics of the composites (the anyons) enters in the form of complex +weights (phases) given in terms of the linking numbers of the worldlines. Hence, the concept of +fractional statistics is intimately related to the theory of knots and of the representations of the +braid group. These concepts were originally introduced in physics to describe statistical trans- +mutation in the theory of solitons [159] in the context of the Skyrme model [160]. Yong-Shi Wu +[161] developed an explicit connection between the work of Leinaas and Myrheim and Wilczek’s +work on anyons in terms of operations acting on the worldlines of the anyons and described by +the Braid Group. Wu’s work and the somewhat early paper by Wilczek and Zee on the statistics +of solitons [162] marked the definite entry of the theory of knots (and of the braid group) in +physics in general and in condensed matter physics in particular. The classification of anyons +in terms of representations of the braid group labeled by the fractional statistics (determined by +linking numbers) as well as of fractional (or topological) spin (determined by the writhing num- +ber of the worldlines) leads to a rich set of physical consequences. As it will turn out, Wilczek’s +anyons are described one-dimensional representations of the braid group. As we will see below, +additional and intriguing (non-abelian) representations will also play a role. +9.2. What is a topological field theory +We will now consider a special class of gauge theories known as topological field theories. +These theories often (but not always) arise as the low energy limit of more complex gauge theo- +ries. In general, one expects that at low energies the phase of a gauge theory be either confining +or deconfined. While confining phases have (from really good reasons!) attracted much atten- +tion, deconfined phases are often regarded as trivial, in the sense that the general expectation +is that their vacuum states be unique and the spectrum of low lying states is either massive or +massless. +Let us consider a gauge theory whose action on a manifold M with metric tensor tensor +gµν(x) is +S = +� +M +dDx √g L(g, Aµ) +(221) +46 + +At the classical level, the the energy-momentum tensor T µν(x) is the linear response of the action +to an infinitesimal change of the local metric, +T µν(x) ≡ +δS +δgµν(x) +(222) +That a theory is topological means that depends only on the topology of the space in which is +defined and, consequently, it is independent of the local properties that depend on the metric, e.g. +distances, angles, etc. Therefore, at least at the classical level, the energy-momentum tensor of a +topological field theory must vanish identically, +T µν = 0 +(223) +In particular, if the theory is topological, the energy (or Hamiltonian) is also zero. Furthermore, if +the theory is independent of the metric, it is invariant under arbitrary coordinate transformations. +Thus, if the theory is a gauge theory, the expectation values of Wilson loops will be independent +of the size and shape of the loops. Whether or not a theory of this type can be consistently defined +at the quantum level is a subtle problem which we will briefly touch on below. +It turns out that, due to the non-local nature of the observables of a gauge theory, the low +energy regime of a theory in its deconfined phase can have non-trivial global properties. In what +follows, we will say that a gauge theory is topological if all local excitations are massive (and in +fact we will send their mass gaps to infinity). The remaining Hilbert space of states is determined +by global properties of the theory, including the topology of the manifold of their space-time. In +several cases, the effective action of a topological field theory does not depend on the metric of +the space-time, at least at the classical level. In all cases, the observables are non-local objects, +Wilson loops and their generalization. +9.3. Chern-Simons Gauge Theory +Gauge theories play a key role in physics. In 2+1 dimensions it is possible to define a special +gauge theory which is odd under time reversal invariance and parity: Chern-Simons gauge theory. +Originally introduced in Quantum Field Theory in1982 by Stanley Deser, Roman Jackiw and +Stephen Templeton [163, 164], Chern-Simons gauge theory can be defined for any compact Lie +group, as well as an extension of Einstein’s gravity in 2+1 dimensions. In 1989 Edward Witten +[165] showed that Chern-Simons gauge theory computes the expectation values of configurations +of Wilson loops, regarded as the worldlines of heavy particles, in terms of a set of topological +invariants known as the Jones polynomial that classify knots in three dimensions. +We will consider the simplest case, the U(1) Chern-Simons gauge theory. The Chern-Simons +action for a U(1) gauge field Aµ in 2+1 dimension is +S [A] = k +4π +� +Ω +d3x ǫµνλAµ∂νAλ + +� +Ω +d3x JµAµ +(224) +where Jµ is a set of conserved currents (representing the worldlines of a set of heavy particles). +On a closed 3-manifold Ω (e.g. a sphere, a torus, etc) the Chern-Simons action is gauge invariant +provided the parameter k (known as the level) is an integer. If the manifold Ω has a boundary, +the action is not gauge invariant at the boundary. Gauge invariance is restored by additional +boundary degrees of freedom. This structure is general, and not just a feature of the U(1) theory. +Since the Chern-Simons action is first order in derivatives it is not invariant under time re- +versal and under parity (which in 2+1 dimensions is a reflection). These symmetries make this +theory relevant to the description of the fractional quantum Hall effect. In the absence of exter- +nal sources, Jµ = 0, at the classical level the Chern-Simons action is invariant under arbitrary +changes of coordinates. This means that the theory is, at least classically, a topological field +theory. +For a general non-abelian gauge group G the Chern-Simons action becomes +S = k +4π +� +M +d3x tr +� +AdA + 2 +3A ∧ A ∧ A +� +(225) +Here, the cubic term is shorthand for +tr +� +A ∧ A ∧ A +� +≡ tr +� +ǫµνλAµAνAλ� +(226) +for a gauge field Aµ that takes values on the algebra of the gauge group G. +47 + +9.4. BF gauge theory +A closely related (abelian) gauge theory is the so-called BF theory [166] which, in a general +spacetime dimension D (even or odd), is a theory of a vector field Aµ (a one-form) and an an- +tisymmetric tensor field B with D − 2 Lorentz indices (a D2 form), known as a Kalb-Ramond +field. In 2+1 dimensions the action of the BF theory is +S = k +2π +� +M +d3x ǫµνλBλ∂µAν +(227) +where, once again, k is an integer. +The BF gauge theory has the same content as the topological sector of a discrete Zk gauge +theory. To see this we will consider the theory of compact electrodynamics which is a U(1) +gauge theory minimally coupled to charged scalar field φ. The will assume that the gauge theory +is defined for a compact U(1) gauge group (meaning that the gauge flux is quantized) and that +the complex scalar field has charge integer k ∈ Z. As usual minimal coupling is implemented by +introducing the covariant derivative Dµ − ∂µ + ikAµ, where Aµ is the U(1) gauge field. Deep in +the phase in which the global U(1) symmetry is spontaneously broken, usually called the Higgs +regime, the amplitude of the scalar field is frozen at its (real) vacuum expectation value φ0 but its +phase ω, representing the Goldstone mode, is unconstrained. In this limit the Lagrangian of this +theory becomes +L = |φ0|2 � +∂µω − kAµ +�2 − 1 +4e2 F2 +µν +(228) +where e is the coupling constant of the gauge field (the electric charge) and Fµν is the field +strength of the gauge field Aµ. In general spacetime dimension D the charge e has units of +length(D−4)/2. In this limit the gauge field becomes massive (this is the Higgs mechanism). This +theory has fluxes quantized in units of 2π/k and has only k distinct fluxes. This is the Zk gauge +theory. +Here we will consider this theory in 2+1 dimensions. We will use a gaussian (Hubbard- +Stratonovich) decoupling of the first term of Eq.(228) in terms of a gauge field Cµ to write the +Lagrangian in the equivalent form +L = − +1 +4|φ0|2C2 +µ + Cµ(∂µω − kAµ) − 1 +4e2 F2 +µν +(229) +Up to an integration by parts, we see that the phase field ω plays the role of a Lagrange multiplier +field which forces the vector field Bµ to obey the constraint ∂µCµ = 0. This constraint is solved +by writing Cµ as +Cµ = 1 +2πǫµνλ∂νBλ +(230) +of a 1-form gauge field Bµ. Upon solving the constraint the Lagrangian of Eq.(229) becomes +L = k +2πǫµνλAµ∂νBλ − 1 +4e2 Fµν(A)2 − +1 +32π2|φ0|2 Fµν(B)2 +(231) +For spacetime dimensions D < 4 the IR the Maxwell terms for the fields Aµ and Bµ are irrelevant +in the IR and, in this limit, this theory reduces to the BF theory at level k of Eq.(227). Therefore, +Zk gauge theory is equivalent to the BF theory at level k. In fact, this result is essentially valid +in all dimensions with the main difference being that the field Bµ is, in general, a rank D − 2 +Kalb-Ramond antisymmetric field. +9.5. Quantization of Abelian Chern-Simons Gauge Theory +By expanding the action of Eq.(224) the Lagrangian density becomes +L = k +4πǫi jAi∂0A j + A0 +� k +2π B − J0 +� +− JiAi +(232) +where B = ǫi j∂iA j is the local flux, J0 is a local classical density and Ji a local classical current. +Then, the first term of Eq.(232) implies that the spatial components of the gauge field obey +equal-time canonical commutation relations +[A1(x), A2(x′)] = i2π +k δ(x − x′) +(233) +48 + +The second term of the Lagrangian enforces the Gauss Law which for this theory simply +implies that the states in the physical Hilbert space obey the constraint +B(x) = 2π +k J0(x) +(234) +Thus the Gauss Law requires that a charge density necessarily has a magnetic flux attached to it. +In other terms, the physical states are charge-flux composites as postulated in Wilczek’s theory +[157]. This is the theoretical basis to the concept of flux attachment which, as we will see in +section 10.3, is widely used in the theory of the fractional quantum Hall effect. +The third term in Eq.(232) simply states that the Hamiltonian density is just +H = JiAi +(235) +Hence, in the absence of sources the Hamiltonian vanishes, H = 0. +The Chern-Simons action is locally gauge-invariant, up to boundary terms. To see this let us +perform a gauge transformation, Aµ → Aµ + ∂µΦ, where Φ(x) is a smooth, twice differentiable +function. Then, +S [Aµ + ∂µΦ] = +� +M +(Aµ + ∂µΦ)ǫµνλ∂ν(Aλ + ∂λΦ) += +� +M +d3x ǫµνλAµ∂νAλ + +� +M +d3x ǫµνλ∂µΦ∂νAλ +(236) +Therefore, the change is +S [Aµ + ∂µΦ] − S [Aµ] = +� +M +d3x ∂µΦF∗ +µ = +� +M +d3x ∂µ(ΦF∗ +µ) − +� +M +d3x Φ∂µF∗ +µ +(237) +where F∗ +µ = ǫµνλ∂νAλ, is the dual field strength. However, in the absence of magnetic monopoles, +this field satisfies the Bianchi identity, ∂µF∗µ = ∂µ(ǫµνλ∂νAλ) = 0. Therefore, using the Gauss +Theorem, we find that the change of the action is a total derivative and integrates to the boundary +δS = +� +M +d3x ∂µ(ΦF∗ +µ) = +� +Σ +dS µΦF∗ +µ +(238) +where Σ = ∂M is the boundary of M. In particular, if Φ is a non-zero constant function on M, +then the change of the action under such a gauge transformation is +δS = Φ × flux(Σ) +(239) +Hence, the action is not invariant if the manifold has a boundary, and the theory must be supplied +with additional degrees of freedom at the boundary. +Indeed, the flat connections, i.e. the solution of the equations of motion, Fµν = 0, are pure +gauge transformations, Aµ = ∂µφ, and have an action that integrates to the boundary. Let the +M = D × R where D is a disk in space and R is time. The boundary manifold is Σ = S 1 × R, +where S 1 is a circle. Thus, in this case, the boundary manifold Σ is isomorphic to a cylinder. The +action of the flat configurations reduces to +S = +� +S 1×R +d2x k +4π∂0φ∂1φ +(240) +This implies that the dynamics on the boundary is that of a scalar field on a circle S 1, and obeys +periodic boundary conditions. +Although classically the theory does not depend on the metric, it is invariant under arbitrary +transformations of the coordinates. However, any gauge fixing condition will automatically break +this large symmetry. For instance, we can specify a gauge condition at the boundary in the form +of a boundary term of the form Lgauge fixing = A2 +1. In this case, the boundary action of the field ϕ +becomes +S [ϕ] = +� +S 1×R +d2x k +4π +� +∂0φ∂1φ − (∂1φ2) +� +(241) +The solutions of the equations of motion of this compactified scalar field have the form φ(x1∓x0), +and are right (left) moving chiral fields depending of the sign of k. This boundary theory is not +topological but is is conformally invariant. +49 + +A similar result is found in non-abelian Chern-Simons gauge theory. In the case of the +S U(N)k Chern-Simons theory on a manifold D × R, where D is a disk whose boundary is Γ, and +R is time, the action is +S CS[A] = +� +D×R +d3x +� k +8πtr +� +ǫµνλAµ∂νAλ + 2 +3ǫµνλAµAνAλ +�� +(242) +This theory integrates to the boundary, Γ×R where it becomes the chiral (right-moving) S U(N)k +Wess-Zumino-Witten model (at level k) at its IR fixed point, λ2 +c = 4π/k +S WZW[g] = +1 +4λ2c +� +Γ×R +d2x tr +� +∂µg∂µg−1� ++ +k +12π +� +B +ǫµνλtr +� +g−1∂µg g−1∂νg g−1∂λg +� +(243) +Here, g ∈ S U(N) parametrizes the flat configurations of the Chern-Simons gauge theory. The +boundary theory is a non-trivial CFT, the chiral Wess-Zumino-Witten CFT [165]. +9.6. Vacuum degeneracy a torus +We will now construct the quantum version of the U(1) Chern-Simons gauge theory on a +manifold M = T 2 × R, where T 2 is a spatial torus, of linear size L1 and L2. Since this manifold +does not have boundaries, the flat connections, ǫi j∂iA j = 0 do not reduce to local gauge transfor- +mations of the form Ai = ∂iΦ. Indeed, the holonomies of the torus T 2, i.e. the Wilson loops on +the two non-contractible cycles of the torus Γ1 and Γ2 are gauge-invariant observables: +� L1 +0 +dx1A1 ≡ ¯a1, +� L2 +0 +dx1A2 ≡ ¯a2 +(244) +where ¯a1 and ¯a2 are time-dependent. Thus, the flat connections now are +A1 = ∂1Φ + ¯a1 +L1 +, +A2 = ∂2Φ + ¯a2 +L2 +(245) +whose action is +S = k +4π +� +dx0ǫi j¯ai∂0¯a j +(246) +Therefore, the global degrees of freedom ¯a1 and ¯a2 at the quantum level become operators that +satisfy the commutation relations +[¯a1, ¯a2] = i2π +k +(247) +We find that the flat connections are described by the quantum mechanics of ¯a1 and ¯a2. A +representation of this algebra is +¯a2 ≡ −i2π +k +∂ +∂¯a1 +(248) +Furthermore, the Wilson loops on the two cycles become +W[Γ1] = exp +� +i +� L1 +0 +A1 +� +≡ ei¯a1, +W[Γ2] = exp +� +i +� L2 +0 +A2 +� +≡ ei¯a2 +(249) +and satisfy the algebra +W[Γ1]W[Γ2] = exp(−i2π/k)W[Γ2]W[Γ1] +(250) +Under large gauge transformations +¯a1 → ¯a1 + 2π, +¯a2 → ¯a2 + 2π +(251) +Therefore, invariance under large gauge transformations on the torus implies that ¯a1 and ¯a2 define +a two-torus target space. +Let us define the unitary operators +U1 = exp(ik¯a2), +U2 = exp(−ik¯a1) +(252) +which satisfy the algebra +U1U2 = exp(i2πk)U2U1 +(253) +50 + +The unitary transformations U1 and U2 act as shift operators on ¯a1 and ¯a2 by 2π, and hence +generate the large gauge transformations. Moreover, the unitary operators U1 and U2 leave the +Wilson loop operators on non-contractible cycles invariant, +U−1 +1 W[Γ1]U1 = W[Γ1], +U−1 +2 W[Γ2]U2 = W[Γ2] +(254) +Let |0⟩ be the eigenstate of W[Γ1] with eigenvalue 1, i.e. W[Γ1]∥0⟩ = |0⟩. The state W[Γ2]|0⟩ is +also an eigenstate of W[Γ1] with eigenvalue exp(−i2π/k), since +W[Γ1]W[Γ2]|0⟩ = ei2π/kW[Γ2]W[Γ1]|0⟩ = e−i2π/kW[γ2|0⟩ +(255) +More generally, since +W[Γ1]W p[Γ2]∥0⟩ = e−i2πp/kW p[Γ2]|0⟩ +(256) +we find that, provided k ∈ Z, there are k linearly independent vacuum states |p⟩ = W p[Γ2]|0⟩, for +the U(1) Chern-Simons gauge theory at level k. It is denoted as the U(1)k Chern-Simons theory. +Therefore the finite-dimensional topological space on a two-torus is k-dimensional. It is trivial +to show that, on a surface of genus g, the degeneracy is kg. +We see that in the abelian U(1)k Chern-Simons theory the Wilson loops must carry k possible +values of the unit charge. This property generalizes to the non-abelian theories, which are tech- +nically more subtle. We will only state some important results. For example, if the gauge group +is SU(2) we expect that the Wilson loops will carry the representation labels of the group SU(2), +i.e. they will be labelled by (j, m), where j = 0, 1 +2, 1, . . . and the 2 j+1 values of m satisfy |m| ≤ j. +However, it turns out SU(2)k Chern-Simons theory has fewer states, and that the values of j are +restricted to the range j = 0, 1 +2, . . . , k +2. +9.7. Fractional Statistics and Braids +Another aspect of the topological nature of Chern-Simons theory is the behavior of expec- +tation values of products of Wilson loop operators. Let us compute the expectation value of +a product of two Wilson loop operators on two positively oriented closed contours γ1 and γ2. +We will do this computation in the abelian Chern-Simons theory U(1)k in 2+1-dimensional Eu- +clidean space. Note that the Euclidean Chern-Simons action is pure imaginary since the action +is first-order in derivatives. The expectation value to be computed is +W[γ1 ∪ γ2] = +� +exp +� +i +� +γ1∪γ2 +dxµAµ +�� +CS +(257) +The result changes depending on whether the loops γ1 and γ2 are linked or unlinked. In this +section we will compute the contribution to this expression for a pair of contours γ1 and γ2. Here +we will not include the contribution to this expectation value for each contours. We will return +to this problem in section 11.3.1 where we discuss the problem of fractional spin. +The expectation value of a Wilson loop on the union of two contours, as in the present case, +γ can be written as +� +exp +� +i +� +γ +dxµAµ +�� +CS = +� +exp +� +i +� +d3xJµAµ +�� +CS +(258) +where the current Jµ is +Jµ(x) = δ(xµ − zµ(t))dzµ +dt +(259) +Here zµ(t) is a parametrization of the contour γ. Therefore, the expectation value of the Wilson +loop is [165] +� +exp +� +i +� +γ +dxµAµ +�� +CS ≡ exp(iI[γ]CS) = exp +� +− i +2 +� +d3x +� +d3y Jµ(x)Gµν(x − y)Jν(y) +� +(260) +where Gµν(x − y) = ⟨Aµ(x)Aν(y)⟩CS is the propagator of the Chern-Simons gauge field. Since the +loops are closed, the current Jµ is conserved, i.e. ∂µJµ = 0, and the effective action I[γ]CS of the +loop γ is gauge-invariant. +The Euclidean propagator of Chern-Simons gauge theory (in the Feynman gauge) is +Gµν(x − y) = 2π +k G0(x − y)ǫµνλ∂λδ(x − y) +(261) +51 + +where G0(x − y) is the propagator of the massless Euclidean scalar field, which satisfies +− ∂2G0(x − y) = δ3(x − y) +(262) +Using these results, we find the following expression for the effective action +I[γ]CS =π +k +� +d3x +� +d3y Jµ(x)Jν(y)G0(x − y)ǫµνλ∂λδ(x − y) +=π +k +� +γ +dxµ +� +γ +dyνǫµνλ∂λG0(x − y) +(263) +Since the current Jµ is conserved, it can be written as the curl of a vector field, Bµ, as +Jµ = ǫµνλ∂νBλ +(264) +In the Lorentz gauge, ∂µBµ = 0, we can write +Bµ = ǫµνλ∂νφλ +(265) +Hence, +Jµ = −∂2φµ +(266) +where +φµ(x) = +� +d3y G0(x − y)Jµ(y) +(267) +Upon substituting this result into the expression for Bµ, we find +Bµ = +� +d3yǫµνλ∂νG0(x − y)Jλ(y) = +� +γ +ǫµνλ∂νG0(x − y)dyλ +(268) +Therefore, the effective action I[γ]CS becomes +I[γ]CS = π +k +� +γ +dxµ +� +γ +dyνǫµνλ∂λG0(x − y) = π +k +� +γ +dxµBµ(x) +(269) +Let Σ be an oriented open surface of the Euclidean three-dimensional space whose boundary is +the oriented loop (or union of loops) γ, i.e. ∂Σ = γ. Then, using Stokes Theorem we write in the +last line of Eq.(269) as +I[γ]CS = π +k +� +Σ +dS µǫµνλ∂νBλ = π +k +� +Σ +dS µJµ +(270) +The integral in the last line of this equation is the flux of the current Jµ through the surface Σ. +Therefore, this integral counts the number of times nγ the Wilson loop on γ pierces the surface Σ +(whose boundary is γ), and therefore it is an integer, nγ ∈ Z. We will call this integer the linking +number (or Gauss invariant) of the configuration of loops. In other words, the expectation value +of the Wilson loop operator is +W[γ]CS = eiπnγ/k +(271) +The linking number is a topological invariant since, being an integer, its value cannot be changed +by smooth deformations of the loops, provided they are not allowed to cross. +We will now see that this property of Wilson loops in Chern-Simons gauge theory leads to the +concept of fractional statistics. Let us consider a scalar matter field that is massive and charged +under the Chern-Simons gauge field. The excitations of this matter field are particles that couple +minimally to the gauge field. Here we will be interested in the case in which these particles are +very heavy. In that limit, we can focus on states that have a few of this particles which will be in +their non-relativistic regime. +Consider, for example, a state with two particles which in the remote past, at time t = −T → +−∞, are located at two points A and B. This initial state will evolve to a final state at time t = T → +∞, in which the particles either go back to their initial locations (the direct process), or to another +one in which they exchange places, A ↔ B. At intermediate times, the particles follow smooth +worldlines. These two processes, direct and exchange. There we see that the direct process is +equivalent to a history with two unlinked loops (the worldlines of the particles), whereas in the +exchange process the two loops form a link. It follows from the preceding discussion that the +52 + +two amplitudes differ by the result of the computation of the Wilson loop expectation value for +the loops γ1 and γ2. Let us call the first amplitude Wdirect and the second Wexchange. The result is +Wexchange = Wdirecte±iπ/k +(272) +where the sign depends on how the two worldlines wind around each other. +An equivalent interpretation of this result is that if Ψ[A, B] is the wave function with the two +particles at locations A and B, the wavefunction where their locations are exchanged is +Ψ[B, A] = e±iπ/kΨ[A, B] +(273) +where the sign depends on whether the exchange is done counterclockwise or clockwise. Clearly, +for k = 1 the wave function is antisymmetric and the particles are fermions, while for k → ∞ +they are bosons. At other values of k the particles obey fractional statistics and are called anyons +[156, 153]. The phase factor φ = ±π/k is called the statistical phase. +Notice that, while for fermions and bosons the statistical phase ϕ = 0, π is uniquely defined +(mod 2π), for other values of k the statistical angle is specified up to a sign that specifies how +the worldlines wind around each other. Indeed, mathematically the exchange process is known +as a braid. Processes in which the worldlines wind clock and counterclockwise are braids that +are inverse of each other. Braids can also be stack sequentially yielding multiples of the phase +ϕ. In addition to stacking braids, Wilson loops can be fused: seen from some distance, a pair of +particles will behave as a new particle with a well defined behavior under braiding. This process +of fusion is closely related to the concept of fusion of primary fields in Conformal Field Theory. +Furthermore, up to regularization subtleties [167], the self-linking terms (those with a = b) +yield a topological spin 1/2k, consistent with the spin-statistics connection [168]. For k = 1 this +means that the flux-charge composites have spin 1/2. +What we have just described is a mathematical structure called the Braid group. The example +that we worked out using abelian Chern-Simons theory yields one-dimensional representations +of the Braid group with the phase ϕ being the label of the representations. For U(1)k there are k +types of particles (anyons). That these representations are abelian means that, in the general case +of U(1)k, acting on a one-dimensional representation p (defined mod k) with a one-dimensional +representation q (also defined mod k) yields the representation one-dimensional p+q (mod k). We +will denote the operation of fusing these representations (particles!) as [q]mod k × [p]mod k = [q + +p]mod k. These representations are in one-to-one correspondence with the inequivalent charges of +the Wilson loops, and with the vacuum degeneracy of the U(1)k Chern-Simons theory on a torus. +A richer structure arises in the case of the non-abelian Chern-Simons theory at level k [165], +such as SU(2)k. For example, for SU(2)1 the theory has only two representations, both are one- +dimensional, and have statistical angles ϕ = 0, π/2. +However, for SU(2)k, the content is more complex. In the case of SU(2)2 the theory has +a) a trivial representation [0] (the identity, (j, m) = (0, 0)), b) a (spinor) representation [1/2] +((j, m) = (1/2, ±1/2)), and c) a the representation [1] ((j, m) = (1, m), with m = 0, ±1). These +states will fuse obeying the following rules: [0] × [0] = [0], [0] × [1/2] = [1/2], [0] × [1] = [1], +[1/2]×[1/2] = [0]+[1], [1/2]×[1] = [1/2], and [1]×[1] = [0] (note the truncation of the fusion +process!). +Of particular interest is the case [1/2] × [1/2] = [0] + [1]. In this case we have two fusion +channels, labeled by [0] and [1]. The braiding operations now will act on a two-dimensional +Hilbert space and are represented by 2 × 2 matrices. This is an example of a non-abelian repre- +sentation of the braid group. These rather abstract concepts have found a physical manifestation +in the physics of the fractional quantum Hall fluids, whose excitations are vortices that carry +fractional charge and anyon (braid) fractional statistics. +Why this is interesting can be seen by considering a Chern-Simons gauge theory with four +quasi-static Wilson loops. For instance in the case of the SU(2)2 Chern-Simons theory the Wilson +loops carry the spinor representation, [1/2]. If we call the four particles A, B, C and D, we +would expect that their quantum state would be completely determined by the coordinates of +the particles. This, however, is not the case since, if we fuse A with B, the result is either a +state [0] or a state [1]. Thus, if the particles were prepared originally in some state, braiding +(and fusion) will lead to a linear superposition of the two states. This braiding process defines +a unitary matrix, a representation of the Braid Group. The same is true with the other particles. +However, it turns out that for four particles there are only two linearly independent states. This +two-fold degenerate Hilbert space of topological origin is called a topological qubit. +53 + +Moreover, if we consider a system with N (even) number of such particles, the dimension of +the topologically protected Hilbert space is 2 +N +2 −1. Hence, for large N, the entropy per particle +grows as 1 +2 ln 2 = ln +√ +2. Therefore the qubit is not an “internal” degree of freedom of the +particles but a collective state of topological origin. Interestingly, there are physical systems, +known as non-abelian fractional quantum Hall fluids that embody this physics and are accessible +to experiments! For these reasons, the non-abelian case has been proposed as a realization of a +topological qubit [169, 170]. +10. Topological Phases of Matter +10.1. Topological Insulators +We will now give a brief discussion of the physics of Topological Insulators from a field- +theoretic perspective. Topological insulators are systems whose electronic states (band struc- +tures) have special topological properties which manifest in the existence of symmetry-protected +edge states. For this reason these systems are known as symmetry-protected topological states +(or SPTs). The simplest example is found in one space dimension where it is related to the fas- +cinating problem of fractionally charged solitons and electron fractionalization. In several ways +many of the concepts involved in these 1+1-dimensional systems can and have been extended to +higher dimensions. +10.1.1. Dirac fermions in 2+1 dimensions +We will now see that, in spite of the formal similarities withe the 1+1 dimensional case, this +theory has different symmetries, particularly concerning parity and time reversal invariance. In +addition, in 2+1 dimensions it is not possible to define a γ5 Dirac matrix which implies that there +is no chiral symmetry and no chiral anomaly. We will consider first a theory of a Dirac field +which in 2+1 dimensions is a bi-spinor (as is in 1+1 dimensions). The Lagrangian of the Dirac +theory coupled to a background gauge field Aµ is +L = ¯ψiγµ∂µψ − m ¯ψψ − eAµ ¯ψγµψ ≡ ¯ψiγµDµ(A)ψ − m ¯ψψ +(274) +where ¯ψ = ψ†γ0, ¯ψγµψ ≡ jµ is the gauge-invariant (and conserved) Dirac current, and Dµ(A) = +∂µ + ieAµ is the covariant derivative. +In Eq.(274), γµ are the three 2 × 2 Dirac matrices which obey the algebra, {γµ, γν} = 2gµν, +where gµν = diag(1, −1, −1) is the metric of 2+1 dimensional Minkowski spacetime. The Dirac +matrices γµ can be written in terms of the three 2 × 2 Pauli matrices. For instance we can choose +γ0 = σ3, γ1 = iσ2 and γ2 = −iσ1 which satisfy the Dirac algebra. Since the three gamma +matrices involve all tree Pauli matrices, it is not possible to define a γ5 matrix which would +anticommute with the gamma matrices. In this sense, it is not possible to define a chirality for +bi-spinors in 2+1 dimensions +We could have also chosen a different set of gamma matrices. For instance, we could have +chosen γ0 = σ3, γ1 = iσ2 and γ2 = +iσ1 which also satisfy the Dirac algebra. These two choices +are equivalent to the 2D parity transformation x1 → x1 and x2 → −x2. In other words, we can +choose the three gamma matrices to be defined in terms of a right handed or a left handed frame +(or triad). Thus, the choice of handedness of the frame used to define the gamma matrices can +be regarded as a chirality. +The parity transformation is also equivalent to a unitary transformation (a change of basis) +U = γ1 = iσ1 which flips the sign of both γ2 and γ0. Thus, a parity transformation is equivalent +to a change of the sign of γ0 or, what is the same, as changing the sign of the Dirac mass +m → −m. Since the Dirac theory is charge conjugation invariant, the parity transformation is +equivalent to time reversal. It is easy to see that in 2+1 dimensions the massive Dirac theory +is not invariant under time reversal since this the single particle Dirac Hamiltonian H(p) for +momentum p involves all three Pauli matrices, +H(p) = α · p + βm +(275) +where, as usual, α = γ0γ and β = γ0. In fact, time reversal is equivalent to the change of the +sign of the Dirac mass. These considerations also apply to the massless Dirac theory coupled to +a background gauge field. +54 + +10.1.2. Dirac Fermions and Topological Insulators +In addition systems in one space dimension, discussed in section 7.1, in which Dirac fermions +naturally describe the low energy electronic degrees of freedom, Dirac fermions also play a role +in many other system in Condensed Matter Physics. Such systems range from the nodal Bogoli- +ubov fermionic excitations of d-wave superconductors in 2D and p-wave superfluids in 3D, to +Chern and Z2 topological insulators in 2D and Z2 topological insulators and Weyl semimetals in +3D. The also play a role in spin liquid phases of frustrated spin-1/2 quantum antiferromagnets, +such as Dirac and chiral spin liquids. Dirac fermions play also a significant role in our under- +standing of the compressible limit of 2D electron gases (2DEG) in large magnetic fields (see +section 10.3). We will not cover here all of these examples. Instead we will focus of the 2D +Chern topological insulators (which exhibit the anomalous quantum Hell effect) and on the 3D +Z2 topological insulators. +In Condensed matter Physics the important low energy electronic degrees of freedom belong +to the electronic states close to the Fermi energy. In such systems the lattice periodic potential +determines the properties of their band structures. The only significant exception to this general +rule is the case of the 2DEGs in GaAs-AlAs heterostructures (the most commonly used platform +for the study of quantum hall effects) whose electronic densities are low enough so that the lattice +periodic potential can for practical purposes almost always be ignored. +In general, Dirac (and Weyl) fermions arise when the conduction and valence bands cross +at isolated points of the Brillouin zone. In such situations, the near the crossing points the low +energy electronic states locally (in momentum space) look like cones and, ignoring possible +anisotropies, look like the states of a Dirac-Weyl theory. For these reasons, Dirac fermions play +a central role in the theory of graphene type materials [171] and, more significantly, in the theory +of topological insulators [172]. +Dirac fermions play a central role in Quantum Electrodynamics and in the Standard Model of +Particle Physics. The problem of quark confinement requires an understanding of these theories +in a regime inaccessible to Feynman-diagrammatic perturbation theory and an intrinsically non- +perturbative formulation, known as Lattice Gauge Theory [55, 56, 173], needed to be developed. +For this reason in the 1970s fermionic Hamiltonians with crossings at points became of great +interest in High Energy Physics as a way to describing the short-distance dynamics of quarks in +Lattice Gauge Theory. +This program run into difficulties when extended to the theory of Weak Interactions which +have an odd number of species of chiral (Weyl) Dirac fermions. All local discretizations of the +Dirac equation yield an even number of species. This fact came to be known as the fermion +doubling problem. These results are special cases of a general theorem, due to Holger Nielsen +and Masao Ninomiya [174, 175, 176] (and extended by Daniel Friedan [177]), which proves that +for systems whose kinetic energy is local in space (as it must be) there must always be an even +number of crossings. Therefore, it is impossible to write a local theory with an odd number of +chiral fermionic species. +In the context of Condensed Matter Physics, the simplest example of Dirac fermions is found +in the electronic structure of graphene (a single layer of graphite) discussed by A. H. Castro Neto +and coworkers [171]. Graphene is an allotrope of crystalline carbon in which the carbon atoms +are arranged in a 2D hexagonal lattice, which is a lattice with two inequivalent sites in its unit +cell, labeled by A and B. Let rA and rB be the sites of the two sublattices. Each site rA has three +nearest neighbors on the B sublattice located at ri +B = rA + di, where i = 1, 2, 3 and +d1 = +� 1 +2 +√ +3 +, 1 +2 +� +, +d2 = +� 1 +2 +√ +3 +, −1 +2 +� +, +d3 = +� +− 1√ +3 +, 0 +� +(276) +(in units in which the lattice spacing is a = 1). The nearest neighbor A sites are related by the +vectors +a1 = + +√ +3 +2 , −1 +2 + , +a2 = (0, 1), +a3 = +− +√ +3 +2 , −1 +2 + +(277) +where a1 and a2 are the primitive lattice vector of the hexagonal lattice. The nearest neighbor +sites of the B sublattice are also related by these same three vectors. +The low energy electronic degrees of freedom of graphene can be described by a single +fermionic state (ignoring spin) at each carbon atom . Let c†(rA) and c†(rB) be the fermion oper- +ators that creates an electron at the site rA and at the site nearest neighbor sites rB, respectively. +The Hamiltonian is +H = t1 +� +rA,i=1,2,3 +c†(rA)c(rA + di) + h.c. +(278) +55 + +where t1 is the hopping matrix element between nearest neighbor A and B sites. +In Fourier space +c(rA) = +� +BZ +d2k +(2π)2ψA(k) exp(ik · rA), +c(rB) = +� +BZ +d2k +(2π)2 ψB(k) exp(ik · rB) +(279) +where BZ is the first Brillouin zone of the hexagonal lattice, a hexagonal region of reciprocal +space with vertices at ±2π/ +√ +3(1, 1/ +√ +3) (which are denoted by K and K′, respectively) and their +two images under 2π/3 rotations. In momentum space the Hamiltonian is [178] +H = t1 +� +BZ +d2k +(2π)2 +� +ψA(k) +ψB(k) +� � +0 +� +j=1,2,3 exp(ik · d j) +� +j=1,2,3 exp(−ik · d j) +0 +� �ψA(k) +ψB(k) +� +(280) +The one-particle states of this Hamiltonian have eigenvalues E(k) = ± +���� +� +j=1,2,3 exp(ik · d j +����. +Clearly E(k) vanishes at the K and K′ points, the corners of the Brillouin zone. Hence, there +is a crossing of the two bands at the K and K′ points near which the energy eigenvalues are a two +cones at K and K′, respectively. Thus, the low energy states of graphene consists of two massless +(gapless) Dirac bi-spinors, with opposite chirality/parity. +Other examples with similar fermion content are a theory of fermions on a 2D square lattice +with a π magnetic flux (1/2 of the flux quantum) per plaquette which arises, for instance, in the +theory of the integer quantum Hall effect on lattices by David Thouless, Mahito Kohmoto, Peter +Nightingale and Marcel den Nijs [179] (albeit for more general flux per plaquette), and in the +theory of flux phases of 2D spin liquids of Ian Affleck and J. Brad Marston [180]. It also arises +in the theory of the chiral spin liquid of Xiao-Gang Wen, Frank Wilczek and Anthony Zee [181]. +We will discuss these theories in section 10. +10.1.3. Chern invariants +The single-particle quantum states of free fermions (electrons) in the periodic potential of a +crystal are Bloch wave functions labeled by the lattice momentum k and band indices m. The +Bloch states are periodic functions of the lattice momenta whose periods are the Brillouin zones. +Topologically each Brillouin zone is a torus: in 1D is a 1-torus, in 2D a 2-torus, etc. The sym- +metries (and shapes) of the Brillouin zone are dictated by the symmetries of the host crystal. +In this section we will consider the quantum states of systems of fermions on a 2D lattice with +M bands with eigenvalues {Em(k)} with m = 1, . . ., M. The eigenstates are Bloch states {|um(k)⟩} +such that the wave functions ar ψm(x) = um(k) exp(ik · x), where k is a (quasi) momentum in the +first Brillouin zone [182]. We will assume that the band spectrum is such that all the bands are +separated from each other by a finite gap for all momenta in the first Brillouin zone. Hence, the +eigenvalues obey the inequality |Em(k) − En(k)| > 0. We will assume that the system of interest +has N < M filled bands and, consequently, that there is a gap separating the top-most occupied +band N (the “valence band”) and the lowest unoccupied band N + 1 (the “conduction band”) that +does not close everywhere in the first Brillouin zone. +Let us consider specifically a 2D system and a Bloch state |um(k)⟩ at a momentum k, and let +∂k|um(k)⟩ be an infinitesimally close Bloch state with momentum k′ = k+dk. The inner product +of these two infinitesimally close Bloch states is the vector field (defined on the first Brillouin +zone for each band m) +A(m) +j (k) = i⟨um(k)|∂ j|um(k)⟩ +(281) +where j = 1, 2 are the orthogonal direction in momentum space and we have used the notation +∂i = ∂/∂ki. This vector field is known as the Berry Connection of the electronic states in the mth +band. +The theory of the quantum states of non-interacting electrons in periodic potentials as devel- +oped by Felix Bloch [182] is the foundation of much of what we know about the physics of many +semiconductors and simple metals. This theory is the core subject of many textbooks [183, 184]. +This theory is based on the classification of the electronic states as representations of the group +of lattice translations and point (and spatial) group symmetry transformations. A key unstated +assumption in much of this body of work is that the Bloch states are globally well-defined func- +tions on the Brillouin zone of a given crystal. It is rather remarkable that this assumption was not +stated explicitly since the original work by Bloch in 1929 until the work by Thouless, Kohmoto, +Nightingale and den Nijs on the quantum states of electrons on 2D lattices in the presence of an +uniform magnetic field [179]. The realization that there is a topological obstruction to define the +56 + +electronic states globally in the Brillouin zone is the key to the development of the the theory of +topological insulators and, more generally, of the modern theory of band structures [185, 186]. +In quantum mechanics states are defined up to a phase. This means that Bloch states that +differ by a phase describe the same quantum state. In other words each quantum state |um(k)⟩ is +a member of a ray (or fiber) of states each labeled by a phase, +|um)(k)⟩ �→ exp(ifm(k)) |um(k)⟩ +(282) +Changing the states by phase factors defines a U(1)M unitary transformation of physically equiva- +lent states. Since we have a fibre at each point k, this amount at defining a fibre bundle. However, +under this transformation the Berry connection A(m) +j (k) changes by a gradient of the phases +A(m) +j (k) �→ A(m) +j (k) + ∂ j fm(k) +(283) +where we assumed that the phases fm(k) are continuous and differentiable functions on the entire +first Brillouin zone. Since the physical states cannot change by redefinitions of the phases of +the basis states we are led to the condition that only the data that is invariant under the gauge +transformations defined by Eqs. (282) and (283) is physically meaningful, which is encoded in +the gauge-invariant pseudo-scalar quantity F (m)(k) +F (m)(k) = ǫi j∂iA(m) +j (k) +(284) +which is known as the Berry curvature. +In what follows we will be interested in the quantity +C(m) +1 += 1 +2π +� +Γ +dk jA(m) +j (k) = 1 +2π +� +BZ +d2k F (m)(k) +(285) +where Γ is the boundary of the 1st Brillouin zone (which we denote by BZ). We will now show +that the quantity C(m) +1 , which measure the flux of the Berry connection A(m) +j (k) through the first +Brillouin zone (in units of 2π), is an integer independent of the particular connection A(m) +j (k) that +we have chosen. This integer-valued quantity is a topological invariant known as the first Chern +number. +However, since the first Brillouin zone is a 2-torus ,which is a closed manifold, a Berry +connection with a net flux cannot obey periodic boundary conditions. Instead, we must allow +for generalized (large) gauge transformations that wrap around the 2-torus of the first Brillouin +zone. This problem is similar (and closely related to) the problem of the wave functions of a +charged particle moving in the presence of a magnetic monopole [151]. We will consider a 2D +system whose first Brillouin zone is spanned by two reciprocal lattice vectors b1 and b2, related +by the two primitive lattice vectors a1 and a2 by the relation bi · a j = 2πδi j (with i, j = 1, 2) The +generalized gauge transformations now are +|um(k + b1)⟩ = exp(if (1) +m (k)) |um(k)⟩, +|um(k + b2)⟩ = exp(if (2) +m (k)) |um(k)⟩ +A(m) +j (k + b1) =A(m) +j (k) + ∂ j f (1) +m (k), +A(m) +j (k + b2) = A(m) +j (k) + ∂ j f (2) +m (k) +(286) +where f (1) +m (k) and f (2) +m (k) are smooth functions of k. +We will now show that the Bloch states cannot be defined globally over the Brillouin zone if +the flux of the Berry connection does not vanish. More specifically let us assume that at some +point k0 ∈ T we know the Bloch state |um(k0)⟩ which satisfies the generalized periodic boundary +conditions of Eq.(286). The question is, can we determine the Bloch state at some other also +arbitrary point k′ +0? We will now that there is a topological obstruction that does not allow it if +the flux of the Berry phase is not zero. The reason is that the phase of the Bloch state cannot be +defined since, in general, the Bloch state will vanish at some point of the Brillouin zone where +the phase is undefined. +We will prove that this si true by following an elegant construction due to Mahito Kohmoto +which goes as follows [187]. The Brillouin zone is a torus, that we will denote by T, can be +regarded as the tensor product of two circles, T ≡ S 1 × S 1. Let us split the torus T into two +disjoint subsets (or patches) HI and HII such that T = HI +� HII. We will assume that in region +HI. Let us assume that the Bloch state vanishes at some point k0 ∈ TI and that the Bloch state +does not vanish for all points k ∈ HII. This means that we can choose the Bloch state |um(k)⟩ +to be real for all k ∈ HII. On the other hand we can always assign some arbitrary phase to the +57 + +Bloch state at k0 ∈ TI. Once we have done that we can extend the definition of the phase to a +neighborhood of k0 wholly included in HI. If we the assume that there is only one zero, then +the phase can be defined over all of HI. The result is that we now have two different definitions +of the phase of the Bloch state on HI and on HII. Let |um(k)⟩I and |um(k)⟩II be the two resulting +definitions of the Bloch state. Let the closed curve γ be the common boundary of regions HI +and HII, γ = ∂HI = ∂HII. However on the common boundary γ these the two definitions of the +Bloch state must be a gauge transformation, +|um(k)⟩I = exp(ifm(k)) |um(k)⟩II +(287) +on all points k ∈ γ. The gauge transformation fm(k), known as the transition function, is a +smooth periodic function of the points k on the closed curve γ. Likewise, the Berry connection +A(m) +j (k) has two definitions on the regions HI and HII which differ by a gauge transformation on +all the points k of the common boundary γ, +A(m) +j (k) +����I − A(m) +j (k) +����II = ∂ j fm(k) +(288) +The transition function defines a mapping of the closed curve γ, which is topologically equivalent +to a circle S 1, to the phase of the Bloch state which is defined mod 2π and hence is also a circle +S 1. Hence the transition functions fm(k) are homotopies which can be classified by the homotopy +group Π1(S 1) ≃ Z. We recognize that the classification of the the transition functions is the same +that we used for the vortices of section 5.1.1. Hence, the change of the transition function on a +full revolution of the closed curve γ must be an integer multiple of 2π. +We will now compute the quantity C(m) +1 , given in Eq.(285), which measures the flux of the +Berry connection through the 2-torus T that defines the first Brillouin zone. Using the partition +of the torus T = TI +� TII we can write +C(m) +1 += 1 +2π +� +T +d2k F (m)(k) = 1 +2π +�� +HI +d2k F (m)(k) + +� +HII +d2k F (m)(k) +� +(289) +Using Stokes Theorem on the two regions HI and HII we get +C(m) +1 += 1 +2π +� +γ +dk j +� +A j(k) +����I − A j(k) +����II +� +(290) +Since the two definitions of the Berry connection differ by a gauge transformation, Eq.(288), we +can express C(m) +1 +in terms of the transition function fm(k) on the curve γ +C(m) +1 += 1 +2π +� +γ +dk · ∂fm(k) = 1 +2π (∆fm(k))γ = n +(291) +where we used that the transition functions are classified by the integer-valued quantity n ∈ Z. +We conclude that the flux of the Berry curvature C(m) +1 +takes only integer values. It is known of +the first Chern number C(m) +1 +which is a topological invariant since it cannot change by a smooth +redefinition of the curve γ. Since C(m) +1 +� 0 implies that the Bloch states must vanish at least +at some point (or points) k ∈ HI, then the integer-valued Chern number can only change if a +redefinition of the curve γ crosses at least one point k at which the Bloch state vanishes. +The conclusion of this analysis is that whenever the flux of the Berry curvature through the +Brillouin zone does not vanish the Bloch states cannot be defined globally on the Brillouin zone. +A direct consequence is that in this case the band is characterized by a topological invariant +called the first Chern number. This number is a property of a given band and, in general, it is +different for each band. This is a particular case of a more general topological classification of +the states of free fermions on a lattice which depends on on the dimensionality of the system +[188, 189, 190]. +10.1.4. The quantum Hall effect on a lattice +We will now show that 2D free fermion systems with Chern bands, i.e. bands characterized +by a non-vanishing Chern number, are insulators that have a quantized Hall conductivity. We will +do this for the theory of the integer quantum Hall effect on a 2D lattice of Thouless, Kohmoto, +Nightingale and den Nijs (TKNN) [179, 187] (see, also, Ref. [191]). Although this problem +played a key role in the development of the theory of topological phases of matter, for many +58 + +years it was viewed as an academic problem since it would require gigantic magnetic fields to +be in the regime of interest for typical solids. The situation changed with the development of +twisted bilayer graphene and similar systems which have unit cells large enough for this physics +to be observable. These new materials has allowed to study this problem experimentally. +TKNN considered a system of free fermions on a planar (actually square) lattice in a uniform +magnetic field B with flux 2πp/q per plaquette (in units in which the flux quantum φ0 = eℏ/c = 1) +with p and q two co-prime integers. The tight-binding Hamiltonian for this simple problem is +H = +� +r, j=1,2 +t j c†(r) eiA j(r) c(r + e j) + h.c. +(292) +where c(r) and c†(r) are fermion creation and annihilation operators on the lattices sites {r = +(m, n)}, e j are lattice unit vectors, and t j, with j = x, y, is a hopping matrix element between +nearest neighbor sites of the square lattice along the x and y directions. On each link of the +lattice we defined a vector potential A j(r), where the oriented sum of the lattice vector potentials +on each plaquette 2πφ, where φ = p +q, is the flux on each plaquette of the square lattice. In the +axial (Landau) gauge A1(m, n) = 0, A2(m, n) = 2πmφ is a periodic function of m with period q. +In this gauge the Hamiltonian becomes +H =t +� +m,n +� +c†(m + 1, n)c(m, n) + c†(m, n)c(m + 1, n) +� ++t +� +m,n +� +c†(m, n + 1) ei2πφm c(m, n) + c†(m, n) e−i2πφm c(m, n + 1) +� +(293) +In this gauge the problem reduces to a system with a theory with a q × 1 unit cell with q inequiv- +alent sites. In momentum space, the (magnetic) Brillouin zone of this system is − π +q ≤ k1 ≤ π +q and +−π ≤ k2 ≤ π. The Fourier transform of the operators c†(r) and c(r) are, respectively, c†(k) and +c(k) which obey the standard anticommutation relations, {c(k), c(q)} = {c†(k), c†(q)} = 0, and +{c†(k), c(q)} = δ(k − q). In momentum space the Hamiltonian becomes +H = q +� π/q +−π/q +dkx +2π +� π +−π +dky +2π +�H(kx, ky) +(294) +where +�H(kx, ky) =1 +q +q−1 +� +n=0 +� +2tx cos(kx + 2πφn) c†(kx + 2πφn, ky) c(kx + 2πφn, ky) ++ ty +� +e−iky c†(kx + 2π(n + 1)φ, ky) c(kx + 2πnφ, ky) + eiky c†(kx + 2π(n − 1)φ, ky) c(kx + 2πnφ, ky) +�� +(295) +The spectrum of this system was first investigated by Hofstadter [192] consists of q bands which +for q an odd integer are separated by finite energy gaps. +For fixed values of kx and ky (in the magnetic Brillouin zone), the Hamiltonian �H(kx, ky) is the +tight-binding model of a one-dimensional chain on the 1D lattice of q sites located at kx + 2πnφ +with nearest neighbor hopping. The single-particle (Bloch) states un(k) of this 1D model obeys +the Schr¨odinger Equation +2tx cos(kx + 2πnφ)un(k) + ty +� +e−iky un−1(k) + eiky un+1(k) +� += Enun +(296) +which is known as Harper’s equation. In general this equation does not admit an analytic solution. +However, the nature qualitative features of the spectrum can be obtained by an expansion either +on tx/ty or on ty/tx. TKNN used degenerate perturbation theory to show that the spectrum has q +bands and that the r-th band is characterized by two integers sr and ℓr which are the solution of +the Diophantine equation +r = q sr + p ℓr +(297) +with ℓ0 = s0 = 0. It isl also obvious that for r = q sq = 1 and ℓq = 0 (for all p). +Furthermore, TKNN showed that there was an, until then unsuspected, relation between the +Hall conductivity σxy of a gapped system of electrons in periodic potentials in a uniform magnetic +field and the Berry connection of the filled bands. This relation implies the quantization of the +59 + +Hall conductivity and its computation in terms of a topological invariant, the Chern number of the +occupied bands. They showed that in this context each band has a non-trivial Berry connection +A(r) +j = i⟨ur(k)|∂ j|ur(k⟩ (with ∂ j = ∂kj) of the form of Eq.(281) with a non-vanishing (first) Chern +number C(r) +1 , i.e. the flux through the magnetic Brillouin zone. +The conductivity tensor characterizes the electrical properties of a physical system. Linear +Response Theory provides a framework for computing the conductivity tensor by perturbing the +system with a weak external electromagnetic field Aµ and computing the currents that they induce +[193, 10]. The expectation value of the gauge-invariant current operator Jµ(x) is +⟨Jµ(x)⟩ = − i +ℏ +δ +δAµ(x) ln Z[Aµ] +(298) +where Z[Aµ] is the partition function in the presence of a background (i.e. classical) electro- +magnetic field Aµ(x). In general spacetime dimension D = d + 1, the lowest order in the vector +potential Aµ, the induced current is given in terms of the polarization tensor Πµν(x, y), +ln Z[Aµ] = i +2 +� +dDx +� +dDy Aµ(x) Πµν(x, y) Aν(y) + O(A3) +(299) +Therefore, the induced current is related to the external field by +⟨Jµ(x)⟩ ≡ Jµ(x) = +� +dDy Πµν(x, y) Aν(y) +(300) +Expressions of this type are know as a Kubo formula. The polarization tensor can be regarded as +a generalized susceptibility. +Although we are using a continuum relativistic notation, these expressions are generally +valid, even in lattice systems. Gauge invariance requires that the polarization tensor be con- +served, +∂µΠµν(x, y) = 0 +(301) +In general, the (retarded) polarization tensor Πµν(x, y) is related to the the (retarded) current- +current correlation function DR +µν(x, y), +Dµν(x, y) = − i +ℏΘ(x0 − y0)⟨[Jµ(x), Jν(y)]⟩ +(302) +by the identity +ΠR +µν(x, y) = DR +µν(x, y) − iℏ +�δJµ(x) +δAν(y) +� +(303) +The last term in Eq.(303), usually called a “contact term”. This term vanishes only for a theory +of relativistic fermions (in the continuum). In all other cases, lattice or continuum, relativistic +or not, the contact term does not vanish and its form depends on the specific theory. In non- +relativistic systems, e.g. in a Fermi liquid, this term is the origin of the f-sum rule [193, 194]. +When the external field Aµ represents is a (locally) uniform electric field E, the induced +current is Ji = σi jE j, where σi j is the conductivity tensor, which can be obtained from the +polarization tensor as the limit +σi j = lim +ω→0 +1 +iω lim +q→0 Πi j(ω, q) +(304) +In a metal, which has a Fermi surface, the order of limits in which ω and q vanish matters and +only the order of limits specified above is the correct one to take. Ina Dirac systems, that we +will discuss below, the order does not matter due to relativistic invariance. The other case in +which the order does not matter is the Chern insulator that we are interested in. In general, in an +isotropic system, the conductivity tensor has a symmetric part and an antisymmetric part. The +symmetric part of the conductivity tensor yields the longitudinal conductivity which has all the +effects of dissipation. The antisymmetric part does not vanish in the presence of a magnetic field +or, more generally, if time reversal symmetry is broken, and yields the Hall conductivity. +A Chern insulator is an insulator and as such is a state with an energy gap. In such a state the +longitudinal conductivity vanishes since there is no dissipation in a gapped state. But in a Chern +insulator time reversal invariance is broken. In the system that we are discussing is broken by +60 + +the magnetic field. The Hall conductivity can be calculated from the antisymmetric part of the +polarization tensor as the limit +σxy = lim +ω→0 +i +ωΠxy(ω, 0) +(305) +In addition, the contact term does not contribute to the Hall conductivity. As a result the hall +conductivity can be computed in terms of the antisymmetric component of the current-correlation +function. +let us now return to the theory of free fermions on a lattice in a (commensurate) magnetic +field takes. As we saw the electronic states are split into q bands with single particle states +|ψn(k)⟩ where k takes values on the first magnetic Brillouin zone. We will compute the Hall +conductivity for this system assuming that the Fermi energy EF lies in the gap between the n-th +and the n − 1-th bands. Let us label by α the occupied bands and by β the unoccupied bands. +Hence Eα(k) < EF < Eβ(k). In this case, the Kubo formula for the Hall conductivity becomes +σxy = ie2 +ℏ +� +Eα 0 and negative for m < 0 or, alternatively for positive and negative chirality of the Dirac +(bi-spinor) field which fixes the sign of the breaking of time reversal invariance. +In addition to the sign, the prefactor of the Chern-Simons term is equal to 1 +2, and it is not +an integer. In section 9.6 we showed that for a dynamical Chern-Simons gauge theory to be +defined consistently on a closed surface this coefficient must be quantized and should take integer +values. The fact that the coefficient is half-quantized means that the classical global symmetry of +a theory of a single Dirac bi-spinor cannot be gauged, which means that the gauge theory cannot +be quantized. This is an example of an obstruction to the quantization of a gauge theory due to +an anomaly [96]. +We will now use these results to compute the effective low-energy action for the theories +of lattice Dirac fermions of section 10.1.2. In those systems the low energy theory is that of +two Dirac (bi-spinors) with different masses m1 and m2. Since both Dirac fields are coupled +(minimally) to the same background gauge field Aµ, the effective Lagrangian is just the sum of +the contribution for each Dirac fermion +Leff[Aµ] = − +1 +4π|m|eff +FµνFµν + 1 +2 +�sgn(m1) + sgn(m2)� 1 +4πǫµνλAµ∂νAλ +(324) +where |m|eff is +1 +|m|eff += +1 +|m1| + +1 +|m2| +(325) +This result implies that, as anticipated in section 10.1.2, this theory has two phases: a parity even +phase and a parity-odd phase. In the regime in which the signs of the two mass terms are equal +and opposite, sgn(m1) = −sgn(m2), the coefficient of the Chern-Simons term cancels and the +low-energy effective Lagrangian is a Maxwell term (generally with an effective speed of light +much smaller than that in vacuum): this phase is a conventional insulator. +Conversely, in the phase in which the two masses have the same sign, sgn(m1) = sgn(m2), +the coefficient of the Chern-Simons terms does not cancel and it is given by ± 1 +4π, where the sign +is the sign of both mass terms. This phase is also an insulator but one in which time-reversal +invariance is broken. Moreover, in this phase the induced current jµ by the background field Aµ +in the long-distance limit is controlled by the Chern-Simons term and it is given by +jµ = δLeff +δAµ(x) = ± e2 +2πℏǫµνλ∂νAλ +(326) +From this result we see that in the parity-broken anomalous quantum Hall phase the system has +a correctly quantized and non-vanishing Hall conductivity +σxy = ±e2 +h +(327) +Therefore, the phase with sgn(m1) = sgn(m2) displays the quantum anomalous Hall effect with a +Hall conductivity whose sign equals the signs of both masses. Notice that this is true regardless +the magnitudes of the masses m1 and m2, and that only their signs matter. In section 10.1 we will +identify this phase with a topological phase of matter known as the Chern Insulator. +We showed that the Hall conductivity of the anomalous quantum Hall phase is, as expected, +equal to e2/h using the effective low energy Dirac theory. One may wonder if this approximation +64 + +may be missing some contributions to the Hall conductivity. Fortunately there is an alternative +quite elegant way of computing the Hall conductivity as a property of the entire occupied band. +This approach, originally introduced by Gregory Volovik [204] in the context of superfluid 3He- +A, and adapted to the theory of the quantum anomalous Hall effect by Viktor Yakovenko [205], +by Maarten Golterman, Karl Jansen and David Kaplan for a theory of Wilson fermions in odd +dimensional hypercubic lattices [206], and by Xiao-Liang Qi, Yong-Shi Wu and Shou-Cheng +Zhang [207], involves the derivation of a topological invariant for a two-band model. +As we saw above the Hall conductivity is obtained from the xy component of the polarization +operator as the limit +σxy = lim +ω→0 +i +ω lim +q→0 Πxy(ω, q) +(328) +For a free fermion system Πxy(ω, 0) is given by the current correlator +Πxy(ω, 0) = +� +BZ +d2k +(2π)2 +� ∞ +−∞ +dΩ +2π tr +� +Jx(k)G(k, ω + Ω)Jy(k)G(k, Ω) +� +(329) +where G(k, Ω) is the propagator for a two-band free fermion system. A generic two-band free +fermion system has a one-particle Hamiltonian of the form given in Eq.(311). The (one-body) +current operator for such a system is obtained as +Jl(k) = ∂h0(k) +∂kl +1 + ∂ha(k) +∂kl +σa +(330) +The propagator G(k, ω) is the 2 × 2 matrix (in band indices) +G(k, ω) = +1 +ω1 − h(k) · σ + iǫ = +P+(k) +ω − E+(k) + iǫ + +P−(k) +ω − E−(k) + iǫ +(331) +where P±(k) are the operators that project onto the (empty) conduction band and the (filled) +valence band whose energies are, respectively E±(k), +P±(k) = 1 +2 +� +1 ± ˆh(k) · σ +� +(332) +where ˆh(k) is the unit vector defined for every momentum k of the first Brillouin zone +ˆh(k) = +h(k) +||h(k)|| +(333) +Upon performing the frequency integration and the band traces in the expression for Πxy(ω, 0) of +Eq.(329), we find that the Hall conductivity takes the form +σxy = e2 +2ℏ +� +BZ +d2k +(2π)2 ǫabc +∂ˆha(k) +∂kx +��ˆhb(k) +∂ky +ˆhc(k) (n+(k)) − n−(k)) +(334) +where n±(k) are the Fermi functions (at zero temperature) for the two bands. Since +E+(k) − E−(k) = 2||h(k)|| = 2 +� +h2(k) > 0 +(335) +there is a non-vanishing gap on the entire Brillouin zone between the occupied valence band, +with n−(k) = 1, and the unoccupied conduction band, with n+(k) = 0. For the insulating state +the Fermi energy lies inside this gap and the expression for the Hall conductivity reduces to the +following +σxy = − e2 +2ℏ +� +BZ +d2k +(2π)2 ǫabc +∂ˆha(k) +∂kx +∂ˆhb(k) +∂ky +ˆhc(k) +(336) +In our discussion of quantum antiferromagnets in one space dimensions in section 5.2 we found +that their effective low-energy action contained a crucial topological term proportional to the +integer-valued topological invariant Q (the winding number) of Eq.(103) which classifies the +smooths maps (homotopies) of the 2D surface (say a sphere S 2) onto the target space of a three- +component unit vector field which also a sphere S 2. These equivalence classes are represented +by the notation Π2(S 2) ≃ Z. In the case at hand the unit vector ˆh(k) are points on a 2-sphere. +Hence, ˆh(k) is a map of the first Brillouin zone (which is a 2-torus) to the sphere S 2. Such maps +65 + +are also classified by the same integer-valued topological invariant defined in Eq.(103). These +results imply that the Hall conductivity of the two-band system is +σxy = e2 +2πℏQ[ˆh] +(337) +In other words we have shown that the Hall conductivity is given in terms of a topological in- +variant of the occupied band (in units of e2/h), the winding number Q. In the two-band model +the topological invariant Q plays the same role as the Chern number does in the work of TKNN +[179, 187] When Q � 0, the two-band system exhibits the quantized anomalous quantum Hall +effect. This si a property of the entire band of occupied states, and not just a consequence of the +low energy approximation. This result implies that the low energy approximation captures all of +the topology of the band. It also implies that in these lattice models the Berry curvature is highly +concentrated near the points in momentum space where the two bands are close in energy. +. +We will now consider the problem of the quantum phase transition between the trivial and +the Chern insulator. To address this problem we will tune the parameters of the lattice model +discussed in section 10.1.2, the phase φ and the ratio of hopping amplitudes t2/t1, to the point at +which the mass of one of the two species of Dirac fermions, say the fermion ψ1, is zero, m1 = 0, +while keeping the fermion ψ2 massive, m2 � 0. In the low energy regime we have a massless +fermion and a massive fermion. This point in parameter space is a quantum phase transition +between a trivial insulator and a Chern insulator. In the low energy regime the fermion ψ1 is +massless. A massless fermion is a scale-invariant system in the sense that the correlators of all +its observables exhibit power law behavior (free field in this case). +We will now discuss briefly the electromagnetic response of this system at the quantum phase +transition where one fermion becomes massless. In particular, it is natural to ask if the coeffi- +cient of the parity-odd (Chern-Simons) term non-vanishing at the quantum phase transition. To +answer this question we will look at the behavior of the parity-even kernel Π0(p2) and the parity- +odd kernel ΠA(p2) in the massless limit for the light fermion, m1 → 0. We find that the total +contribution of both the light fermion and of the heavy fermion to the polarization kernels is +(assuming m2/m1 → ∞) +lim +m1→0 Π0(p2) = +i +16 +� +p2 , +lim +|m2|→∞ ΠA(p2) = − 1 +4πsgn(m2) +(338) +The important conclusion is that at this quantum critical point the parity-even kernel Π0(p2) +is non-local and that the heavy regulator fermion (the “doubler”) yields the leading finite non- +vanishing and local contribution to the parity-odd kernel ΠA(p2). +We can now use these results to compute the conductivity tensor σi j at the quantum critical +point where m1 → 0. Since the system is spatially isotropic the conductivity tensor has the form +σi j = +� σxx +σxy +−σxy +σxx +� +(339) +where we used that σyy = σxx since the system is isotropic. The longitudinal conductivity σxx +and the Hall conductivity σxy are +σxx = π +8 +e2 +h , +σxy = ±1 +2 +e2 +h +(340) +in other words, at the quantum critical point the system has a finite (and universal) longitudinal +conductivity. This result may seem surprising as there is no disorder in this model. A finite +universal longitudinal conductivity is a standard occurrence in 2D systems at a quantum critical +point. For example, at the superconductor-insulatortransition the conductivity is (conjectured) to +be) σxx = e2/2h (with σxy = 0). In addition, it also has finite and also universal Hall conductivity. +The Hall conductivity at the quantum critical point is due to the heavy fermionic “doubler” and is +equal to 1/2 (in units of e2/h). So, the quantum critical point is not time-reversal invariant since +this symmetry is broken at the UV (lattice). +We can examine the massless theory using a more formal approach [199, 208]. In a gauge- +invariant regularization of the massless theory the partition function of a Dirac fermion coupled to +a background (unquantized) U(1) gauge field, the partition function is not time-reversal invariant. +66 + +It is given by +Z[Aµ] = +� +D ¯ψDψ exp +� +i +� +d3x ¯ψi /D[Aµ]ψ +� += det(i /D[Aµ]) = +����det(i /D[Aµ]) +���� exp +� +±iπ +2η[Aµ] +� +(341) +where the sign depends on how time-reversal invariance is broken by the choice of regularization. +The quantity η[Aµ] that appears in the phase factor is the Atiyah-Patodi-Singer η-invariant [147] +which we already encountered in the discussion of the fractionally charged solitons in section 8.3. +With some caveats [199, 208], the phase factor of the partition function of Eq.(341) is commonly +written in the form of a 1/2-quantized Chern-Simons term +π +2η[Aµ] ≡ ±1 +2 +� +d3x 1 +4πǫµνλAµ∂νAλ +(342) +often denoted as a U(1)1/2 Chern-Simons term. This term plays the same role as the contribution +of the heavy fermion doubler in the lattice theory. +However, we can also wonder if there is a way to have a time-reversal invariant theory of a +single massless Dirac fermion, m → 0. This case cannot be realized in a y 2D lattice model but, +as will see, it can be realized on the surface states of a 3D time-reversal invariant Z2 topological +insulator, which we will discuss in section 10.1. +10.2. Three-dimensional Z2 topological insulators +The classification of topological insulators in terms of a Chern number is only possible in +even space dimensions in systems with broken time reversal symmetry: in d = 2 the Berry +connection is abelian and the topological invariant is the first Chern number while in in d = 4 the +Berry connection in a non-abelian SU(2) gauge field and the topological invariant is the second +Chern number [135], etc. We will now discuss the time-reversal invariant topological insulators +[209, 190, 135] and, in particular, those that are invariant under inversion symmetry. Such states +exist in both two and space dimensional insulator with strong spin-orbit coupling. +10.2.1. Z2 Topological Invariants +Let {|un(k)⟩} be the Bloch states. We will represent time reversal by the anti-unitary operator +Θ that acts on the single particle (Bloch) states by complex conjugating the state and reverses the +spin. For spin-1/2 fermions Θ = exp(iπσ2)K, where K is the complex conjugation operator. In +this case Θ2 = −1. Let us assume that we have two occupied Bloch bands for each point k f the +Brillouin zone. In this case the states form a rank-2 vector bundle over the torus of the Brillouin +zone. In time reversal invariant systems the anti-unitary time reversal transformation T induces +an involution in the Brillouin zone that identifies the points k and −k. Time reversal the acts +on the one-particle (Bloch) Hamiltonian as ΘH(k)Θ−1 = H(−k). The states |un(±k)⟩ are related +by time reversal as |un(−k)⟩ = Θ|un(k)⟩ which implies that the bundle is real. The condition +Θ2 = −1 implies that the bundle is real. In algebraic topology these bundles are classified by an +integer (here the number of occupied bands) and a Z2 index that will allow us to classify these +states. +In a periodic lattice there exists a set of points Qi of the Brillouin zone with the property +that they differ by their images under the action of time reversal by a reciprocal lattice vector, +−Qi = Qi + G. In d = 2 there four such points and d = 3 there are eight points, and are given +by Qi = 1 +2 +� +j n jb j, where n j = 0, 1, j = 1, 2 in d = 2 and j = 1, 2, 3 in d = 3. Here b j are +the primitive lattice vectors. Kane and Mele [210] defined the 2N × 2N antisymmetric matrix +�m,n(k) = ⟨um(−k)|Θ|un(k)⟩. They showed that at each time reversal invariant point Qi one can +define an index δi +δi = +� +det[�(Qi)] +Pf[�(Qi)] += ±1 +(343) +where det[�] and Pf[�] are, respectively, the determinant and the Pfaffian of the matrix �, and +det[�] = Pf[�]2. The sign of the quantities δi can be made unambiguous by requiring that the +Bloch states be continuous. In addition, the quantities δi are gauge-dependent. However the +products +(−1)ν = +4 +� +i=1 +δi +(344) +67 + +in d = 2, and +(−1)ν0 = +8 +� +i=1 +δi, +(−1)νk = +� +nk=1,nj�k=0,1 +δi(n1, n2, n3) +(345) +are gauge and are also topological invariant. The Z2-valued indices ν and ν0 are robust to disorder +and are called strong topological indices. Furthermore, Fu and Kane showed that if ξn(Qi) = ±1 +are the parity eigenvalues of the occupied parity eigenstates, the quantities δi are given by δi = +�N +m=1 ξ2m(Qi). In d = 3 the index ν0 does not rely on the existence of inversion symmetry. +In the case of a two-band model in d = 2 and in d = 3, the states are four-component spinors +reflecting the two bands and the two spin components. In the context of these systems with +strong spin-orbit coupling spin is actually the z-component of the atomic total angular momentum +J of the electrons with energies close to the Fermi energy. In these systems the one-particle +Hamiltonian H(k) is a 4 × 4 Hermitian matrix which can be expanded as a linear combination +of Dirac matrices. A simple and very useful model of systems of this type is the Wilson fermion +model (with continuous time) [55, 211, 135] of a square (cubic) lattice with 4 states per site +(parity and spin) whose Hamiltonian three dimensions is +H(k) = sin k · α + M(k) β +(346) +where α = γ0γ and β = γ0 are the conventional 4 × 4 Dirac matrices. The γ matrices satisfy +the Clifford algebra {γµ, γν} = 2gµν1, where gµν = diag(1, −1, −1, −1) is the metric tensor of +four-dimensional Minkowski space time. An additional γ matrix of interest is γ5 = iγ0γ1γ2γ3. +The (Wilson) mass term M(k) in two dimensions is M(k) = M + cos k1 + cos p2 − 2, and in +three dimensions is M(k) = M + cos k1 + cos k2 + cos k3 − 3. In Eq.(346) we see that, consistent +with the requirements of the Nielsen-Ninomiya theorem [174, 175], in addition to a possible low +energy Dirac fermion (if M is small) there are three more massive Dirac fermions (in d = 2) and +seven other Dirac fermions in d = 3, so that the total number of Dirac fermions is always even, 8 +in this case. With Wilson’s mass term the additional Dirac fermions (the “doublers”) are always +heavy even if the Dirac fermion near the Γ point Q = 0 is light. +In the Dirac basis time reversal is the operation Θ = (iσ2 ⊗ I)K, where K is complex con- +jugation, and parity is P = β. The matrices α and β commute with PΘ. At the time-reversal +and parity invariant points of the Brillouin zone {Qi} the Hamiltonian only depends of the the +matrix β, H(Qi) = M(Qi)β. Since the parities of the spinors are the eigenvalues of the matrix +β, we conclude that in the two-band models the quantities δi are simply equal to the sign of +the mass of the fermions defined at the time-reversal-invariant points Qi of the BZ [209, 135], +δi = δ(Qi) = −sgn M(Qi). Using this result it follows that in two dimensions the system is a Z2 +topological insulator with index ν = 1 (mod 2) if 0 < M < 4 while it is trivial for other values +of M, i.e. ν = 0 mod 2. Similarly, in three dimensions a Z2 (strong) topological insulator exists +only if 0 < M < 2 (in the other regimes this system is in a weak topological insulator state or in +a trivial one). +We conclude that both in two and three dimensions the low energy theory of the Z2 topolog- +ical insulators consists of a single Dirac fermion whose mass is small compared to that of the +fermion doublers and has the opposite sign. In both cases there is a quantum phase transition +between a trivial insulator and the time-reversal invariant topological insulator at the point where +he mass of the light fermion vanishes. +10.2.2. The Axial Anomaly and the Effective Action +We will now look at the electromagnetic response of a Z2 topological insulator in 3+1 dimen- +sions. This problem can be addressed more easily using the continuum field theory description +which is valid in the regime where the mass M is weak. In this regime one species of Dirac +fermions is light (i.e. its mass is small) while the Dirac doublers remain heavy. Much as we did +in our discussion of the Chern insulators and the parity anomaly in section 10.1.6 we will keep +in mind that the fermion doublers play the role of heavy regulators, such as in the Pauli-Villars +scheme, in Quantum Field Theory. Here too, we will see that a field-theoretic anomaly, known +as the axial anomaly plays a central role in the physics. The analysis is very similar to what we +did with the chiral anomaly in 1+1 dimensions in section 7.3. +Let us begin with a theory of a single massive Dirac fermion in 3+1 dimensions. The La- +grangian of the free massive Dirac theory is +L = ¯ψ � i/∂ − m� ψ +(347) +68 + +The equation of motion of the spinor field operator ψ(x) (I omit the spinor indices here) is the +Dirac equation +�i/∂ − m� ψ = 0 +(348) +The Dirac Lagrangian has a global gauge symmetry ψ → eiθψ which requires that the Dirac +current jµ = ¯ψγµψ is locally conserved, ∂µ jµ = 0. The massless Dirac theory can be decomposed +into a theory of two Weyl bi-spinor fields which obey separate Dirac Lagrangians. Furthermore, +the massless theory has the additional global symmetry under the transformation ψ → eiθγ5ψ and +the additional formally locally conserved axial current j5 +µ = i ¯ψγµγ5ψ. However the conservation +law of the axial current is violated in the massive Dirac theory +∂µ j5 +µ = −2mi ¯ψγ5ψ +(349) +since the two Weyl bi-spinors transmute into each other in the presence of a mass term. This is +the origin of the phenomenon of neutrino oscillations. In the condensed matter physics context +there is a similar phenomenon in systems of Weyl semimetals which have crossings between the +valence and conducting bands at two locations ±Q of the BZ, with each crossing associated with +each Weyl bi-spinor. A charge density wave with ordering wavevector 2Q mixes the two Weyl +fermions which become gapped, becoming effectively a single massive Dirac fermion [212]. +This state is often called an “axionic”-charge-density-wave. +We will now show that the axial symmetry has an anomaly and cannot be gauged. Thus, we +will consider the problem of a Dirac theory coupled to a background U(1) gauge field Aµ and +reexamine the putative conservation of the axial current j5 +µ. This question can be addressed in +different ways. Quite early on Steven Adler [133], and John Bell and Roman Jackiw [134] exam- +ined this problem by computing a Dirac fermion triangle diagram for the process of a neutral pion +decaying into two photons, π0 → 2γ. In particle physics the pion is the Goldstone boson of the +spontaneously broken chiral symmetry. The analog of this problem in condensed matter physics +is the phase mode of an incommensurate charge density wave. The relativistic Lagrangian for +this problem is a theory of Dirac fermions coupled to a complex scalar field φ = φ1 +iφ2 through +two Yukawa couplings +L = ¯ψi /Dψ + gφ1 ¯ψψ + igφ2 ¯ψγ5ψ − V(φ1, φ2) = ¯ψi /Dψ + g|φ| ¯ψeiγ5θψ − V(|φ|2) +(350) +where |φ|2 = φ2 +1 + φ2 +2, tan θ = φ2/φ1, and Dµ = ∂µ + ieAµ is the covariant derivative. Here we are +regarding the gauge field Aµ as a background probe field. +The triangle Feynman diagram computes the polarization tensor for the electromagnetic field +Aµ with an insertion of the coupling to the complex scalar field in an otherwise massless theory. +Assuming a gauge-invariant regularization of the diagram, this computation finds that the axial +current j5 +µ is anomalous and is not conserved even in the massless theory [133, 134] +∂µ j5 +µ = − e2 +16π2 FµνF∗ +µν +(351) +where F∗ +µν = 1 +2ǫµνλρFλρ is the dual of the electromagnetic field tensor. This is the axial anomaly. +To see how the axial anomaly arises we will follow the physically transparent approach of +Nielsen and Ninomiya [176], that we also employed in section 7.3 in 1+1 dimensions. We will +consider a theory of free massless dirac fermions coupled to a background electromagnetic field. +Since the theory is massless, the Dirac equation decouples into an equation to the right handed +Weyl fermion ψR (with positive chirality γ5ψR = +ψR) and a left handed Weyl fermion ψL (with +negative chirality, γ5ψL = −ψL). In the gauge A0 = 0, the Dirac equations become +[i∂0 − (−i∂ − eA) · σ]ψR = 0, +[i∂0 − (i∂ − eA) · σ]ψL = 0 +(352) +Let us now consider look at the solutions of the Weyl equation for right-handed fermions ψR, +Eq.(352). The left-handed fermions ψL are analyzed similarly. Wd will consider a gauge field +A1 = 0 and A2 = Bx1 representing a uniform static magnetic field of strength B pointing along +the x3 direction. In this this gauge the eigenstates are plane waves along the directions x2 and +x3 and harmonic oscillator states along the direction x1. The eigenvalue spectrum consists of +Landau levels with energies +E(n, p3, σ3) = ± +� +2eB +� +n + 1 +2 +� ++ p2 +3 + eBσ3 +�1/2 +(353) +69 + +for n = 0, 1, 2, . . ., except for the zero mode with n = 0 and σ3 = −1, for which +E(n = 0, p3, σ3 = −1) = ±p3 +(354) +where the + sign holds for ψR and the − sign for ψL. Just as in the case of non-relativistic fermions +the relativistic Landau levels are degenerate. We will consider a system of Dirac fermions at +charge neutrality and, hence, EF = 0. The positive and negative energy states are charge conju- +gate of each other and the negative energy states are filled. For right-handed fermions, the zero +mode states with p3 < 0 are filled while for right handed states the zero modes with p3 > 0 are +filled. The density of states of the zero modes is LeB/(4π2) (where L is the linear size of the +system). +Let us consider now turning on an external electric field E parallel to the magnetic field B. +Just as we saw in 1+1 dimensions in section 7.3, the electric field leads to pair creation by shifting +the Fermi momentum to pF for the zero modes. There is no particle creation for the states with +n � 0 and they do not contribute to the anomaly. The rate of creation of right-handed fermions +NR, +dNR +dx0 += 1 +L +LeB +4π2 +pF +dx0 += e2 +4π2 EB +(355) +The annihilation rate of left handed particles is +dNL +dx0 += − 1 +L +LeB +4π2 +pF +dx0 += − e2 +4π2 EB +(356) +and the creation rate of left handed anti-particles is +d ¯NL +dx0 += 1 +L +LeB +4π2 +pF +dx0 += e2 +4π2 EB +(357) +the axial anomaly is the total rate of creation of right handed particles and of left-handed antipar- +ticles: +dQ5 +dx0 += dNR +dx0 ++ d ¯NL +dx0 += e2 +2π2 EB +(358) +which agrees with the expression of Eq.(351). +We now turn to the Z2 topological insulator. As we saw this is a system with two species of +(4 component) Dirac spinors. In the topological phase the sign of the Dirac mass term one of the +Dirac fermions (the one near the Γ point in the lattice model) is opposite (negative) to the sign +of the mass term of the of the other Dirac fermion (the fermion doubler). Explicit calculations +[135, 213, 214] on the lattice model obtain the result that the effective low-energy action for the +electromagnetic gauge field Aµ in the topological phase is +S eff[Aµ] = +� +d4x +� +− 1 +4e2 FµνFµν + +θ +32π2e2ǫµνλρFµνFλρ +� ++ . . . +(359) +In a time-reversal invariant system the allowed values of the θ angle of Eq.(359) are restricted to +be θ = nπ, with n ∈ Z. The case θ = 0 (mod 2π) represents a trivial insulator whereas θ = π (mod +2π) holds for a Z2 time-reversal invariant topological insulator. +The second term in the effective action of the electromagnetic gauge field of Eq.(359), known +as the θ term, has been extensively discussed in the high-energyphysics literature [202, 215, 216]. +The derivation of this term is subtle. As it stands, unless θ is varying in space-time (in which case +this is known a the axion field) this term is a total derivative. In non-abelian Yang-Mills gauge +theories this term is proportional to a topological invariant known as the Pontryagin index which +counts the instanton number of the gauge field configurations [99, 93]. In the context of the +Lagrangian of Eq.(350) this term is induced by the coupling of the Dirac fermion to the Yukawa +coupling of the complex scalar field φ to the Dirac and γ5 mass terms. In this case, the lowest +order contribution is given by the triangle diagram. The fact that this term is exact at lowest order +reflects the fact that the axial anomaly is in fact a non-perturbative effect which is the same in +both the weak coupling and the strong coupling regimes [96]. +In the phase where the potential V(|φ|2) in the Lagrangian of Eq.(171) has a minimum at +|φ0| exp(iθ) the chiral symmetry is spontaneously broken and the phase field θ(x) is the Goldstone +boson of the spontaneously broken chiral symmetry. In this phase the Dirac fermions become +massive through the Yukawa couplings to the complex scalar field, and the phase of the field φ +enters in the effective action as an axion field which couples to the gauge field Aµ through the θ +term of the effective action. We should note that in a recently studied axionic CDW state of a +Weyl semimetal [212] the phase of the CDW plays the role of the axion field. +70 + +10.2.3. Theta terms, and Domain walls: Anomaly and the Callan-Harvey Effect +We will now discuss some remarkable behaviors of three-dimensional Z2 topological in- +sulators. We will begin with the electromagnetic response encoded in the effective action of +Eq.(359). We will assume that the θ angle is effectively a slowly varying Goldstone mode of the +spontaneously broken U(1) chiral symmetry (i.e. the axion field) present in the Lagrangian of +Eq.(350). If the field θ is constant, the θ term is a total derivative and it does not contribute to +the local equations of motion. We will see below that this term [lays a key role in the physics of +a domain wall, which we will regard as the interface of a Z2 topological insulator and a trivial +insulator. In the general case in which θ varies slowly (as Goldstone modes do) its presence leads +to interesting modification of Maxwell’s equations, known as axion electrodynamics [215]: +▽ · E =˜ρ − e2 ▽ θ · B, +▽ × E = − ∂tB +▽ × B =∂tE + ˜j + e2 (∂tθ + ▽θ × E) , +▽ · B =0 +(360) +where ˜ρ and ˜j are external probe electric charge and current densities. The equations of axion +electrodynamics have many remarkable properties. Here we will focus on effects: the topological +magnetoelectric effect [135] and the Witten effect [217]. +Let us consider a Z2 topological insulator with a flat open boundary (the x1 − x2 plane) +perpendicular to the direction x3. We will assume that the topological insulator lies at x3 < 0. +This means that for x3 < 0, and far from the surface, θ(x3) → π, while in the trivial vacuum, and +also far from the surface, θ(x3) → 0. We will assume that the change of θ from 0 to π occurs +on a short distance ξ. We will call this configuration an axion domain wall. In the region where +θ(x3) is changing, |x3| ≲ ξ, an applied uniform magnnetic field B induces a uniform electric +field E parallel to B whose magnitude is proportional to the change in θ. This is the topological +magnetoelectric effect [135]. +A similar striking effect is obtained by considering the case of a magnetic monopole of mag- +netic charge 2π/e (as required by Dirac quantization) inside a sphere of the trivial region of radius +R, with θ = 0, surrounded by a region with θ � 0. The two regions are separated by a thin (axion) +wall in which θ changes for 0 to π in a narrow shell of thickness ξ ≪ R. The equations of axion +electrodynamics imply that the magnetic monopole induces an electric charge Qe on the surface +of the sphere +Qe = e∆θ +2π +(361) +Thus, a magnetic monopole acquires an electric charge and becomes a dyon. This is the Wit- +ten effect [217]. In the particular case of a Z2 topological insulator, time reversal invariance +requires that ∆θ = π, which implies that a monopole with unit magnetic charge has an elec- +tric charge e/2. A similar argument implies that an external magnetic field perpendicular to the +open surface of the Z2 topological insulator (which is essentially an axion domain wall) induces +an electric charge polarization on the surface proportional to the total magnetic flux, which is +another manifestation of the topological magnetoelectric effect. +The reader should readily recognize that the result of Eq.(361) for the electric charge induced +my the magnetic monopole is the same as the Goldstone-Wilczek equation for the fractional +charge for a one dimensional soliton of Eq.(215). We will now see that this is not just an analogy. +We will follow here the general approach of Curtis Callan and Jeffrey Harvey [218] who extended +the earlier work of Goldstone and Wilczek [141]. Let us consider the bulk of a Z2 topological +insulator and assume that θ(x) is slowly varying. We can use the triangle diagram calculation to +find that an electromagnetic gauge field Aµ induces a current ⟨Jµ(x) given by +⟨Jµ(x)⟩ = −i +e +16π2 ǫµνλρ +φ∗(x)∂νφ(x) − φ(x)∂νφ∗(x) +|φ(x)|2 +Fλρ(x) = +e +8π2 ǫµνλρ∂νθ(x)Fλρ(x) +(362) +This result implies that, in the case of a domain wall in the x1 − x2 plane, a magnetic field +perpendicular to the wall induces a current towards the wall and, hence, a charge accumulation +on the wall. Where does this charge come from? To understand this problem we will consider +a Z2 topological insulator occupying a slab of macroscopic size L between a wall at x3 = 0 and +a far way “anti-wall” at x3 = L. In this configuration a magnetic field normal to the wall(s) +induces a transfer of charge from one wall to the other. Similarly, an electric field parallel to the +wall induces a current also parallel to the wall and perpendicular to the electric field, i.e. a Hall +current. +In 1+1 dimensions we saw that soliton configurations acquire a fractional charge associated +to states bound to the soliton (which are zero modes when ∆θ = π). We will see that the surfaces +71 + +of the 3D Z2 topological insulators also have zero modes which ar Weyl fermions propagating +on the wall. To see how this works we will consider a 3+1 dimensional Dirac fermion with a +Dirac mass that changes sign at x3 = 0. The Lagrangian now ill be +L = ¯ψi/∂ψ + gφ(x) ¯ψψ +(363) +where φ(x) is now a real scalar field that has the asymptotic behaviors φ(x3) = φ0 f(x3) such +that limx3→±∞ f(x3) = ±1. We will assume that f(x3) is a monotonous function of x3, but its +actual dependence on x3 is immaterial aside from the requirement that it should change sign at +some point which we will take to be x3 = 0. This is a special case of Eq.(350) with φ1 = φ and +φ2 = 0. We will now recognize that this just a Dirac fermion with a position-dependent Dirac +mass m(x3) = gφ(x3). The one-particle Dirac Hamiltonian for this system is +H = −iα · ▽ + m(x3)β +(364) +where α and β are the four 4 × 4 Dirac matrices. By symmetry, this Hamiltonian can be split into +two Hamiltonians, H = Hwall + H⊥, +Hwall = −iα1∂1 − iα2∂2, +H⊥ = −iα3∂3 + m(x3)β +(365) +Let the spinor ψ± be and eigenstate of the anti-hermitian Dirac matrix γ3 = βα3 with eigenvalues +±i, respectively +γ3ψ± = ±iψ± +(366) +We seek a spinor solution ψ± of the Dirac equation Eq.(364) such that H⊥ψ± = 0 +± ∂3ψ± + m(x3)ψ± = 0 +(367) +which is also a solution of +(iγ0 − iγ1∂1 − iγ2∂2)ψ± = 0 +(368) +In other words, it is a solution of the massless Dirac equation in 2+1 dimensions. The require- +ment that ψ± be an eigenstate of γ3 reduces the number of spinor components from four to two. +The full solution has the form +ψ± = η±(x0, x1, x2)F±(x3), +±∂3F±(x3) = −m(x3)F±(x3) +(369) +For a domain wall with limx3→∞ m(x3) = +m, the normalizable solution is F+(x3) is +F+(x3) = F(0) exp +� +− +� ∞ +0 +dx′ +3 m(x′ +3) +� +(370) +where F(0) is a constant. For the anti-domain wall, for which limx3→∞ m(x3) = −m, the normal- +izable solution is F−(x3). +We conclude that there is a 2+1-dimensional massless Dirac theory that describes the quan- +tum states bound to the wall which propagate along the wall. These states are the generalization +of the fractionally charged mid-gap states of solitons in one dimension discussed in section 8.3. +The energy of these states is E(p) = ±|p|, where p = (p1, p2) (where I set the Fermi velocity to +unity). Experimental evidence for 2+1 dimensional massless Dirac fermions on the surface of the +3D Z2 topological insulator Bi2Te3 has been found in spin-polarized angle-resolved photoemis- +sion studies of the surface states which showed that they have a linear energy-momentumrelation +(expected of massless Dirac fermions) as well as the spin-momentum locking characteristic of +these spinor states [219]. +Having succeeded in showing that the surface of the Z2 time-reversal invariant 3D topological +insulator has a two-component massless Dirac spinor we now want to determine its electromag- +netic response. In fact we have already discussed this problem in our discussion of the parity +anomaly in section 10.1.6 where we showed that the effective action for the electromagnetic field +Aµ of Eq.(342) of a single massless Dirac bi-spinor is a Chern-Simons term with a prefactor +which is 1/2 of the allowed value. In that purely 2+1-dimensional context we saw that time re- +versal invariance is actually broken. However, the 3D problem is time-reversal invariant so there +must be a contribution that cancels this time-reversal anomaly. The answer is that the requisite +cancellation is supplied by the bulk. +To see how this can happen we return to the bulk effective action of the 3D topological +insulator of Eq.(359) where we observed that the θ-term is a total derivative. Suppose that the +72 + +system has a boundary at x3 = 0 and that the topological insulator exists for x3 > 0 (where the +fermion mass is negative, m < 0) and trivial for x3 < 0 (where m > 0). Only the region with +m < 0 contributes to the θ-term. The region of four-dimensional space time occupied by the +topological insulator is M and in this region the θ-term becomes +S θ[A] = +θ +32π2 +� +M +d4x ǫµνλρFµνFρλ = +θ +8π2 +� +M +d4x ǫµνλρ∂µAν∂λAλ = +θ +8π2 +� +∂M +d3x ǫµνλAµ∂νAλ +(371) +where ∂M ≡ Σ × R is the boundary of the region M, Σ is the surface of the 3D topological +insulator and R is time. +Thus we see that the θ-term of the effective action S θ[A] integrates to the boundary where it +has the form of a 2+1-dimensional Chern-Simons term. Since for a time-reversal-invariant 3D +topological insulator θ = π, we see that in this case the boundary Chern-Simons term becomes +S θ=π[A] = 1 +8π +� +∂M +d3x ǫµνλAµ∂νAλ +(372) +with a coefficient which is 1/2 of the allowed value. This bulk contribution either cancels the +parity anomaly of the boundary state, rendering the full system time-reversal invariant. In other +words, in this system time-reversal symmetry is realized by cancellation of the anomaly between +the bulk of the topological insulator and the boundary or, equivalently, by an inflow of the parity +anomaly between the boundary and the bulk of the system. Another way to phrase this result is +the statement that a single Dirac fermion cannot exist on its own in 2+1 dimensions but it can +as the boundary state of a 3+1-dimensional system whose anomaly cancels the anomaly of the +boundary. +10.3. Chern-Simons Gauge Theory and The Fractional Quantum Hall Effect +We will now turn to the problem of the quantum Hall effects. This is a problem that revealed +the existence of profound and far reaching connections between condensed matter physics, quan- +tum field theory, conformal field theory and topology. In particular, the fractional quantum hall +effect is the best studied and best understood topological phase of matter. As such, it has become +the conceptual springboard for its manifold generalizations. +The integer (IQHE) and fractional (FQHE) quantum Hall effects are fascinating phenomena +observed in fluids of electrons in two dimensions in strong perpendicular magnetic fields. The +integer quantum Hall effect was discovered by Klaus von Klitzing in 1980 [220] in transport +measurements of the longitudinal and Hall resistivity of the surface states of metal oxide field- +effect transistors (MOSFET) in magnetic field of up to 15 Tesla. The effect that von Klitzing +discovered (for which he was awarded the 1985 Nobel Prize in Physics) was that in the high +field regime the measured Hall conductivity showed a series of sharply defined plateaus at which +took the values σxy = ne2/h, where n is an integer, and the longitudinal conductivity appeared +to vanish σxx → 0 as the the temperature was lowered down to T ≃ 1.5 K. Remarkably the +measured value of the Hall conductivity was obtained with a precision of ∼ 10−9. To this date +the measurement of the Hall conductivity in the IQHE yields the most precise definition of the +fine structure constant. +Subsequent transport experiments in ultra-high-purity GaAs-AlAs heterostructures by Dan +Tsui, Horst St¨ormer and Art Gossard found that, in addition to the IQHE, two-dimensional elec- +tron fluids in high magnetic fields exhibit the fractional quantum Hall effect meaning that there +are equally sharply-defined plateaus of the Hall conductivity at the values σxy = p +q +e2 +h , where p +and q are co-prime integers [221]. Much as in the IQHE case, in the FQHE the longitudinal +conductivity vanishes at low temperatures. It is important to note that both in the IQHE and in +the FQHE the observed temperature dependence in the highest purity samples of the longitudinal +conductivity is activated, σxx ∼ exp(−W/T). The observed value of the energy scale W is the +experimental estimate of an energy gap in the electron fluid in the quantum Hall states. +In addition to MOSFETS and GaAs-AlAs heterostructures both the IQHE and the FQHE +have been seen in several other experimental platforms, particularly in graphene and other 2D +materials [222, 223]. The explanation of these effects and of a panoply of startling consequences +that were uncovered in the course of understanding this phenomenon is the focus of this section. +10.3.1. Landau levels and the Integer Hall effect +At some level the integer quantum Hall effect can be explained by the Landau quantization of +the energy levels of a free charged moving in two dimensions in a perpendicular magnetic field +73 + +[224]. The Hamiltonian for a non-relativistic particle of charge −e and mass M in a perpendicular +magnetic field B is +H = +1 +2M +� +−iℏ ▽ +ie +c A(x) +�2 +(373) +for a uniform perpendicular magnetic field B = B ˆez = ▽ × A(x). +In the circular gauge the vector potential is Ai = − 1 +2 Bǫi jx j. We will assume that the 2D plane +has linear size L. The total magnetic flux is Φ = BL2 and we will assume that there is an integer +number Nφ of magnetic flux quanta piercing the plane, φ = Nφφ0, where φ0 = hc/e is the flux +quantum. In units such that ℏ = e = c = 1 the flux quantum is φ0 = 2π and Φ = 2πNφ. +In the presence of a magnetic field the components of the canonical momentum operator +p = −iℏ ▽ − e +c A do not commute with each other, +[pi, p j] − ieℏ +c Bǫi j +(374) +This means that translations in two directions do not commute with each other. However, the +components of the operator k = p(−B) commute with p (and hence with the one-particle Hamil- +tonian) but do not commute with each other: the commutator is the same as in Eq. (374). Since +[k, H] = 0 they act as symmetry generators of the group of magnetic translations. For arbitrary +displacements a and b the translation operators t(a) = exp(ia· k/ℏ) (and similarly with b) satsify +t(a)t(b) = exp(ia × b · ez/ℓ2 +0)t(b)t(a) +(375) +Magnetic translations only commute with each other is the area subtended by a and b contains +an integer number of magnetic flux quanta. +Given the rotational symmetry of the circular gauge it is natural to work in complex coordi- +nates z = x1 + ix2. We will also use the notation ∂z = (∂1 − i∂2)/2 and ∂¯z = (∂1 + i∂2)/2. In this +gauge, up to a normalization, the eigenstate wave functions have the form +ψ(z, ¯z) = f(z, ¯z) exp +− |z|2 +4ℓ2 +0 + +(376) +where ℓ0 = +� +ℏc +e|B| is the magnetic length. In this gauge (and in complex coordinates) the angular +momentum operator Lz = −iℏ(x1∂2 − x2∂1) = ℏ(z∂z − ¯z∂¯z). +Any analytic function f(z) is an eigenstate with energy E0 = +1 +2ℏωc. A complete basis of +analytic functions are the monomials fn(z) = zn and have energy E0 and angular momentum +Lz = nℏ. This is the lowest Landau level whose wave functions are ψn(z) = zn exp(−|z|2/4ℓ2 +0). On +the other hand, an anti-analytic function fN = ¯zN is an eigenstate of energy EN = ℏωc +� +N + 1 +2 +� +, +where ωc = +e|B| +Mc is the cyclotron frequency, and angular momentum Lz = −Nℏ. States with +angular momentum nℏ have the same energy and the degeneracy is equal to the number of flux +quanta Nφ. For the most part we will be interested in the states in the lowest Landau level. +In the absence of disorder the Landau levels have an extensive degeneracy equal to the num- +ber of flux quantum Nφ. If we consider a system of N electrons in a Landau level the natural +measure of density is not the areal density ρ = Ne +L2 but the fraction ν = Ne +Nφ the states in the Landau +level which are occupied by electrons. The many-body state in which all Nφ states of the lowest +Landau level has filling fraction ν = 1. The wave function for this state is the Slater determinant +of the Landau states in the m = 0 level. After some simple algebra the wave function of this state +is found to be +Ψν=1(z1, . . . , zNe) = +� +i< j +(zi − zj) exp +− +Ne +� +i=1 +|zi|2 +4ℓ2 +0 + +(377) +This is the ground state of the non-interacting system and it is non-degenerate. It is easy to +see that this state has a finite energy gap. Indeed, in the Hilbert space with a fixed number +of electrons, the lowest energy excitation is a particle-hole pair in which the particle is in an +unoccupied state of the first excited Landau level, with m = 1 and energy E1 = 3 +2ℏωc and a hole +a single-particle state in the lowest Landau level, with energy E0 = ℏωc. The excitation energy of +the electron-hole pair is just ℏωc which is finite. However, in the free particle system this excited +states has a huge degeneracy as the particle and the hole can be in any single particle state of the +Landau level. +74 + +At a very naive level one can estimate the Hall conductivity for a translationally invariant +system filling up n Landau level by the following simple argument. If n Landau levels are filled, +the number of electrons N = nNφ and the total charge is then Q = eN = e n Nφ = e n BL2/φ0 = +ne2BL2/hc. If a weak and uniform in-plane electric field E is applied in the presence of a per- +pendicular magnetic field B = B ez, then the entire system (its center of mass) moves at the drift +velocity � such that � × B = −Ec. Thus, there is a Hall current J = Q�. The current density +is j = J/L2 = Q�/L2. Hence, ji = +Qc +BL2 ǫi j E j, which implies that the Hall conductivity is +σxy = Qc/BL2 = ne2/h. +While the above argument is formally correct and yields the correct value of the Hall conduc- +tivity in this highly idealized setting, at a very basic level it is faulty. To begin with, it assumes +exact translational invariance and, hence, Galilean symmetry, which are not obeyed in any realis- +tic system. The second objection is that, even in this idealized system, as the number of electrons +is varied, the chemical potential (and hence the Fermi energy) jumps discontinuously from one +Landau level to the next resulting in a linearly increasing Hall conductivity (as predicted classi- +cally) without any of the observed plateaus. Furthermore, this argument does not explain which +the Hall conductivity is so precisely quantized whereas, a priori, one would expect that being a +transport coefficient the Hall conductivity would depend on lots of complicated materials details. +But, it it turns out that it does not! +Let us first discuss the observed universality of the Hall conductivity, namely its robustness +and independence of microscopic details. In a remarkable paper Qian Niu, David Thouless and +Yong-Shi Wu showed that, provided the ground state is separated by a finite energy gap from +the excited states, the Hall conductivity is actually a topological invariant [225]. The argument +has a similarity with what we discussed in the case of the Hall effect on a 2D lattice (in section +10.1.4) but in this more general system one considers the full many-body wave function, rather +than the one-particle states. To show that this true they considered a system of N electrons +with periodic boundary conditions, i.e. on a two-dimensional torus, T 2 = S 1 × S 1, with Nφ +flux quanta going through the torus. The torus is a rectangle of dimensions L1 and L2 with +opposite ends identified. This is necessary to allow for a Hall current to be globally allowed. The +electromagnetic gauge field for a system ona torus with uniform non-vanishing flux cannot obey +periodic boundary conditions but rather accross the torus the gauge fields must differ by (large) +gauge transformations +A1(x1, x2 +L2) = A1(x1, x2)+∂1β2(x1, x2), +A2(x1 +L1, x2) = A2(x1, x2)+∂2β1(x1, x2) (378) +while the wave functions themselves obey twisted boundary conditions +Ψ([x(j) + L1e1]) = exp +−i e +ℏc +Ne +� +j=1 +β1([x(j)]) + iθ1 + × Ψ([x(j)]) +Ψ([x(j) + L2e2]) = exp +−i e +ℏc +Ne +� +j=1 +β2([x(j)]) + iθ2 + × Ψ([x(j)]) +(379) +where e j (with j = 1, 2,) are two unit vectors on the torus, and θ1 and θ2 are two angles that twist +the boundary conditions of the wave functions. +In order to have a current on the torus they assumed that, in addition to the uniform flux going +through the torus, there a weak uniform electric field E on the torus represented by the constant +in space vector potential δA = E t = ▽[U(x)t] whose circulation on the two non-contractible +circles Γ1 and Γ2, which wrap around the directions x1 and x2 of the torus T 2, are +I j = +� +Γj +δA · dx = t +� +Γj +E · dx = itE jL j +(380) +Line integrals of a gauge field on non-contractible loops in space (or space-time) are called +holonomies. By inspection we see that the angles θ1 and θ2 are given by +θj = e +ℏcI j +(381) +Alternatively, the angles θ = (θ1, θ2) can be interpreted as magnetic fluxes (in units of the flux +quantum φ0) through the two non-contractible circles of the torus T 2. +The angles θ are defined mod 2π since they are phase factors that twist the phase of the +wave functions, and define a 2-torus of boundary conditions. By comparing with what we did in +75 + +section 10.1.4 in the case of the single-particle wave functions in the lattice model, we see that +the many-body wave function Ψθ and the twist angles θ is the same as the relation between the +phase of the Hofstadter wave functions with the momentum k of the magnetic BZ (which is also +a 2-torus). Indeed, as it is always the case, while the phase of (in this case) the many-body wave +function Ψθ is arbitrary, the changes of this phase as the the twist angle θ is changed are not. To +quantify this dependence, for a given many-body state Ψ(α) +θ , where α labels the state, we define a +Berry connection A(α)(θ) on the torus of boundary conditions θ +A(α)(θ) = i +� +Ψ(α) +θ +����∂θ +����Ψ(α) +θ +� +(382) +Under a redefinition of the phase of the state +Ψ(α) +θ +→ exp(if(θ)) Ψ(α) +θ +(383) +the Berry connection A(α)(θ) changes by a gauge transformation +A(α)(θ) → A(α)(θ) − ∂θ f(θ) +(384) +Niu, Thouless and Wu [225] showed that the expression of the Kubo formula for the Hall +conductivity of the state Ψ(α) +θ , averaged over the boundary conditions θ, is given in terms of the +flux of the Berry connection A(α)(θ) through the 2-torus of boundary conditions, by the gauge- +invariant quantity +⟨(σxy)α⟩θ = e2 +ℏ +� 2π +0 +dθ1 +2π +� 2π +0 +dθ2 +2π +� +∂1A(α) +2 +− ∂2A(α) +1 +� += e2 +h +1 +2π +� +γ +A(α)(θ) · dθ +(385) +where γ is the square contour with corners at (0, 0), (2π, 0), (0, 2π) and (2π, 2π), that defines the +2-torus of boundary conditions θ. This result is known as the Niu-Thouless-Wu formula. +Eq.(385) implies that for the Hall conductivity to be non-vanishing the Berry connection +A(α)(θ) must have a non-vanishing flux through the torus of boundary conditions. For this to be +true, just as we did in section 10.1.3, we must consider large gauge transformations of the form +A1(θ + 2πe2) =A1(θ) + ∂1 f2(θ) +A2(θ + 2πe1) =A2(θ) + ∂2 f1(θ) +Ψ(α)({x(j)}; θ + 2π e1) = exp(if1(θ)) Ψ(α)({x(j)}; θ) +Ψ(α)({x(j)}; θ + 2π e2) = exp(if2(θ)) Ψ(α)({x(j)}; θ) +(386) +where e1 and e2 are two orthogonal unit vectors on the torus of boundary conditions. We can +now repeat the analysis we did in section 10.1.3 which, in this context, implies that the wave +function Ψ(α) +θ +cannot be globally well defined on the torus of boundary conditions, where it must +have zeros, and where it should be defined in patches. The end result is that the wave functions +labeled by α are classified by a topological invariant, the first-valued Chern number C(α) +1 , which +here it is given by +C(α) +1 += 1 +2π +� +γ +A(α)(θ) · dθ +(387) +and, hence, that the Hall conductivity in he state α (averaged over the twisted boundary condi- +tions) exhibits the integer quantum Hall effect, +⟨(σxy)(α)⟩θ = C1 +e2 +h +(388) +Expressing the Hal conductivity in terms of the first Chern number, which is a topological invari- +ant, proves that the value of the conductivity cannot be changed by local physics effects, such +as disorder, etc. The topological nature of the Hall conductivity is the reason for the robustness +and high precision of the measured value of the Hall conductivity. In subsequent work Niu and +Thouless showed that, provided the state has a finite energy gap to all excitations, in the thermo- +dynamic limit the Hall conductivity averaged over boundary conditions and for a given boundary +condition are equal. +The result of Eq.(388) seemingly implies that there can only be an integer quantum Hall +effect which, as we see shortly, it is not the case: there is a fractional quantum Hall effect in +76 + +which the Hall conductivity is a fraction, σxy = p +q +e2 +h , where p and q are co-prime integers. This +value of the Hall conductivity is also found experimentally to be obeyed with the same precision +as in the integer quantum Hall effect. In other words, the Hall conductivity in the fractional +case must also have a topological character. To understand why this can be the case we need an +implicit assumption that we made in our derivation. In fact, in deriving the result of Eq.(388) we +assumed (implicitly) that for each value of θ there is only one many body state Ψ(α) +θ +(up to gauge +transformations). As we will see below, in the fractional quantum Hall effect on a torus (and, in +fact on any closed surface except the sphere) the ground state is degenerate in the thermodynamic +limit. We will also see that traversing the torus of boundary conditions once maps one degenerate +state to another one. In general, we will find that on the 2-torus in the thermodynamic limit there +are m ∈ Z exactly degenerate states. In this case, one returns to the original state after sweeping +m times the torus of boundary conditions. We will also see below that the degenerate states are +labeled by quantum numbers related to the anyons that they support. +While the topological argument proves the robustness (and universality) of the value of the +Hall conductivity, it does not provide an insight for why there are plateaus in its magnetic field +dependence. If for some value of the filling fraction ν = N/Nφ the electron fluid is exactly at a +ground state Ψ(α) upon changing either the number of particles N or the magnetic field B (but not +both) the fluid will now will be at the state Ψ(α) plus or minus some number of electrons. The +existence of the plateau in σxy can be understood if the extra electrons (or holes) do not contribute +to the Hall conductivity. This happens in the presence of disorder as these extra particles will +be localized by the disorder and localized states, whose wave functions decay exponentially are +insensitive to boundary conditions and, hence, these states do not contribute to the conductivity. +This argument was first put forth by Laughlin [226] who build on the result that in a disordered +system in two dimensions all states are localized [63] but made the key assumption that the +state at the center of the Landau level (broadened by disorder) is extended and contributes to +the conductivity if this state is filled. This picture implies that this is actually a quantum phase +transition from an insulator (dubbed a Hall insulator) to a state with a quantized Hall conductivity. +This intuitive and appealing proposal has been the focus of research for many years and +it is largely an unsolved problem. There is in fact a proposed field theory for this quantum +phase transition based on a a non-linear sigma model (in replica space) with a topological θ-term +[227, 228], analogous to the theory we discussed in section 5.2 for quantum antiferromagnets in +1+1 dimensions. in this proposal the (Boltzmann) longitudinal conductivity σxx plays the role of +the inverse of the coupling constant of the non-linear sigma model while σxy plays the role of the +coefficient of the θ-term. Although the RG flows suggested by this approach qualitatively agree +with currently existing experimental data, and actual derivation is still wanting. +However, aside from the fact that is is still an unsolved problem, there is the problem that this +explanation ignores the fact that in a Landau level interactions are dominant and are the key to +the understanding of the fractional case even in the same samples! A symptom of this problem +is that, at least in the cleaner samples, as we noted above the longitudinal conductivity vanishes +exponentially with temperature. This dependence means that there is a clean energy gap (and not +just a mobility gap) which suggests that the state with additional excitations may have some sort +of at least local order which would provide for a local energy gap. An actual theory based on this +picture has yet to be developed. +10.3.2. The Laughlin Wave Function +We will turn now to the fractional quantum Hall effect (FQHE). The FQHE is observed in +fractionally filled Landau levels. These states have fractional filling fractions, namely that ν = Ne +Nφ +is a fraction. Most of the observed states are in the lowest Landau level, with N = 0, but a few +states with very intriguing properties are seen in the first excited Landau Level with N = 1. +Let us consider for now the states in the lowest Landau level, N = 0. We will assume that +there is no disorder and, for now, we will ignore the effects of boundaries of the sample. Physi- +cally, the important interaction here is the Coulomb interaction although, in some samples with +screening layers, short-range interactions may be important as well. Since there is a large mag- +netic field, the Zeeman interaction is typically a large energy scale and the electros are expected +to be fully polarized. However, there are situations under which the gyromagnetic factor can be +made small (and even zero) and spin unpolarized states have to be considered. In some platforms, +such as graphene, additional quantum numbers are present, such as “flavor” associated with the +different Dirac states. In our discussion we will not cover these richer possibilities. +Under these circumstances we have a problem in which all Nφ available one-particle states are +77 + +exactly degenerate (the Landau level is “flat”) but only a fraction of them are filled. In this limit, +this system as as strongly interaction as it gets. Not surprisingly, in a system of this type time- +honored standard approximations fail, such as Hartree-Fock. Given the macroscopic degeneracy +and the nature of the states in a magnetic field, a system of this type may harbor many different +types of actual ground states depending of the types of interactions that are considered. +Tsui, St¨ormer and Gossard [221] did transport experiments in the two-dimensional electron +gas (2DEG) in high-purity GaAs-AlAs heterostructures and reported the discovery of a state with +a highly precise value of the Hall conductivity of σxy = 1 +3 +e2 +h . Such a state cannot be understood +in any obvious way in terms of the integer quantum Hall effect. The observation of a clean gap +in the low temperature longitudinal resistivity σxx → 0 as T → 0 implied that the 2DEG is +incompressible and quite likely uniform. +The first breakthrough was Laughlin’s unique insight which led to his explanation of the +experiments in terms of a novel quantum liquid of fully spin polarized electrons at filling fractions +ν = 1 +m (with m and odd integer) whose wave function he proposed to be +Ψm(z1, . . ., zN) = +� +i< j +(zi − zj)m × exp +− +Ne +� +i=1 +|zi|2 +4ℓ2 +0 + +(389) +where N = Ne is the number of electrons, {zi} are their (complex) coordinates and m is an odd +integer (which makes the wave function antisymmetric, as required by Fermi statistics). +Laughlin extracted the physics encoded in this wave function by considering the probability +����Ψm(z1, . . ., zN) +���� +2 +of finding the N electrons in the locations {zi} (with i = 1, . . ., N). The norm of +the Laughlin wave function is +���� +����Ψm +���� +���� +2 += +� +d2z1 . . . d2zN +����Ψm(z1, . . ., zN) +���� +2 +(390) +The Laughlin wave function Ψm(z1, . . ., zN) is an eigenstate of angular momentum with eigen- +value Lm = 1 +2mN(N − 1). This remarkable wave function has a large overlap (∼ 98%) with the +exact wave function with V(R) = e2 +εR (Coulomb) interactions in systems of up to eight particles +(ε is the dielectric constant). +The norm of the state can ten be interpreted as the classical partition function of a system +of a system of N particles at the locations {zi} at inverse temperature β = m with the effective +interaction +U(z1, . . ., zN) = −2 +� +1≤i< j≤N +ln |z1 − zj| + +1 +2mℓ2 +0 +N +� +j=1 +|zj|2 +(391) +This classical partition function is a one-component plasma, of a gas of particles with unit (neg- +ative) charge interacting with each other with the repulsive 2D classical (logarithmic!) Coulomb +interaction, VCG = − ln |zi − zj| (not 1/R!). The last term represents the contribution of a neu- +tralizing background positive charge (which here it originates in the gaussian factor of the wave +function). In other words, this is a one-component plasma. We can check that this is correct +since ▽2 +� +|z|2 +2mℓ2 +0 +� += +2 +mℓ2 +0 which corresponds to a uniform (areal) charge density ρ0 = +1 +2πmℓ2 +0 . Clas- +sical Monte Carlo simulations of this probability distribution showed that this wave function +describes an incompressible fluid state if m ≲ 7, while for larger values it represents a crystalline +state. This approach is known as the plasma analogy. +10.3.3. Quasiholes have fractional charge +Laughlin considered a vortex-like excitation in the fluid created by the adiabatic insertion of +an infinitesimally thin solenoid at some coordinate z0 carrying a magnetic flux Φ, a fluxoid. In +the presence of this solenoid the single particle state zn exp(|z|2/4ℓ2 +0) acquires a branch cut and +becomes the state zn+α exp(|z|2/4ℓ2 +0) where α = Φ/φ0, with φ0 = hc/e being the flux quantum. +This analytic structure is the Aharonov-Bohm effect [155]. +However, if the solenoid carries just one flux quantum α = 1 and the net effect is that the state +becomes zn+1 exp(−|z|2/4ℓ2 +0) which has one more unit of angular momentum. For these reasons, +in the presence of a solenoid with one flux quantum inserted at z0, the Laughlin wave function +for the 2DEG in a large magnetic field now becomes +Ψqh +m (z0; z1, . . ., zN) = +N +� +i=1 +(zi − z0) × Ψm(z1, . . . , zN) +(392) +78 + +A calculation using the plasma analogy reveals that the net effect of the solenoid is to expel a +charge equal to −e/m from the vicinity of the location of the solenoid or, equivalently, to create a +fractionally charged quasihole at z0 with charge Qqh = +e/m. The charge distribution is uniform +at long distances but is depleted (hence a quasihole) on a length scale ξ. One can check that this +state costs a finite energy ε0 which depends on the particular interaction. +On the other hand, if we consider a system in which the solenoid is inserted adiabatically +so that after a long time t there is a quasihole at z0, an amount of charge equal to −e/m will +flow radially outwards to the outer edge of the 2DEG. The adiabatic insertion of the solenoid +generates a azimuthal electric field (an e.m.f.) and a radial Hall current, and a Hall conductivity +σxy = 1 +m +e2 +h . Hence, the Laughlin wave function exhibits the fractional quantum Hall effect and +its excitations are vortices with fractional charge e/m. +It is important to note the non-local nature of the Laughlin quasihole. The non-locality is +inherited from the fluxoid (the solenoid) inserted at z0: even though the magnetic field of the +fluxoid vanishes away from it, B(x) = Φ δ2(x − x0), its vector potential A(x) does not since its +circulation on any non-contractible loop γ that contains the location of the fluxoid inside must +be equal to the flux Φ carried by the fluxoid. Although in principle one can choose a singular +gauge in which A(x) vanishes locally, this cannot be true everywhere as it must leave behind a +Dirac-like string on a curve Γ ranging form the location z0 of the fluxoid to the boundary of the +system with A taking a singular value on Γ. This feature has the same form as the 2D vortex that +we discussed in section 5.1.1. Kivelson and Ro˘cek [229] showed that the existence of this Dirac +string causes the phase of a quantum state that carries charge e∗ to have a branch cut on the curve +Γ and to change by exp(i2π(e∗/e)Φ/φ0), as required by the Aharonov-Bohm effect [155]. This +effect is unobservable for integer-charged states, i.e. electrons. +10.3.4. The Jain States +The construction of the Laughlin quasihole motivated Jainendra Jain [230, 231] to rewrite +the Laughlin wave function in the form +Ψm(z1, . . . , zN) = +� +i< j +(zi − zj)m−1 × Ψν=1(z1, . . ., zN) +(393) +Here the prefactor represents particles each attached with an even number, m − 1, of flux quanta. +The second factor,Ψν=1(z1, . . ., zN), is the wave function of fermions filling up a lowest Landau +level, i.e. the Vandermonde determinant of Eq.(377). In this form, the inverse of the filling factor, +1/ν, which is the number of flux quanta per particle, is written as 1 +ν = (m − 1) + 1 = m. The +interpretation now is that the (m − 1)φ0 magnetic flux carried by each particle, which partially +screens the uniform field down to an effective field Beff = B − (m − 1)φ0 corresponding to one +effective flux quantum per particle, i.e. a filled lowest Landau level of the effective field Beff. In +other words, the wave function for the ν = 1/m FQH state is reinterpreted as the νeff = 1 integer +QH state of N composite fermions, each being an electron carrying attached an even integer m−1 +flux quanta. +This procedure is known as flux attachment. Physically it means that the strong interactions +between the electrons for other electrons to be further away which, in virtue of being in a strong +magnetic field, is equivalent to an increase of their relative angular momentum, as if there was a +local change in the magnetic field. There is no reason to believe that the composite fermions are +weakly coupled, even though this is frequently stated in the literature without any real supporting +evidence. +Jain then generalized this reinterpretation of the Laughlin state to a whole sequence of FQH +states obtained by filling p levels of these partially screened magnetic fluxes +Ψm,p(z1, . . . , zN) = PLLL +� +i< j +(zi − zj)m−1Ψν=p(z1, . . ., zN) +(394) +where Ψν=p(z1, . . ., zN) is the wave function of p filled Landau levels (of composite fermions), +and PLL is an operator that projects onto the lowest Landau level. By counting fluxes we see that +the filling fractions for the Jain states satisfy +1 +ν(m, p) = m − 1 ± 1 +p +(395) +79 + +where we allowed for the possibility that the external fluxes may be over-screened by the com- +posite fermions. Alternatively we can write +ν(m, p) = +p +p(m − 1) + ±1 +(396) +The Jain states with p = 1 are the laughlin states and each is called the primary state of a +Jain sequence. Almost all the observed FQH states belong to one of these sequences (and its +generalizations). In particular, the most prominent states (i.e. those with larger plateaus) belong +to the sequence 1 +3, 2 +5, 3 +7, 4 +9, 5 +11, . . . and to the reversed sequence 1, 2 +3, 3 +5, 4 +7, . . .. For p → ∞, the Jain +sequences converge to the values of the filling fraction limp→∞ ν(m, p) = +1 +m−1. In this limit, the +m − 1 fluxes attached to each particle exactly cancels th external flux leading us to a (possibly) +Fermi liquid of composite fermions [232]. +10.3.5. Quasiholes have fractional statistics +The non-locality and topological nature of the Laughlin quasihole implies that this state must +be regarded as a soliton (or vortex) of the charged quantum fluid. The vortex-like state with wave +function Ψqh +m (z0; z1, . . . , zN) is called the Laughlin quasihole. This state represents a composite +object of a flux quantum and a fractional charge. This structure of this quantum state is strongly +reminiscent of the flux-charge composite objects considered by Frank Wilczek who showed that +such objects should be anyons, states that exhibit fractional statistics [156]. Although at a con- +ceptual level anyons were proposed (almost) prior to the discovery of the FQHE, it is in this +setting that fractional statistics entered actual physics. The actual experimental observation took +work by many people, and was only confirmed in 2020 [233]. +Fractional statistics dictates the analytic form of the wave functions of two or more quasiholes +[234]. Let us consider a laughlin-type wave function for two quasiholes located at (complex) +coordinates u and �. Naively we expect the wave function for two quasiholes in the ν = 1/m +Laughlin state to have the form +Ψ(u, �; z1, . . . , zN) = N(u, �) +N +� +j=1 +(zj − u)(zj − �) Ψm(z1, . . ., zN) +(397) +The prefactor N(u, �) must be chosen to account for the fact that theres is an additional change of +angular momentum due to the additional fluxoid at �. We also expect that as the quasihole with +charge e/m is adiabatically carried around the the other quasihole (also with charge e/m) there +will be ac accrued phase in the wave function due to the Aharonov-Bohm effect of the charge +of one quasihole circling around the flux of the other (or, equivalently, crossing the branch cut) +[229]. The requirement of translation invariance and analyticity are met by the choice [234] +(ignoring a multiplicative normalization constant) +N(u, �) = (u − �)1/m exp +− +1 +4ℓ2 +0m(|u|2 + |�|2) + +(398) +which has a branch cut stretching from u to �. A calculation using the plasma analogy represents +this state as a set of N charge −1 particles interacting with two additional particles of charge −1/m +at u and � (and a neutralizing background). The branch cut implies that dragging a quasihole +around the other during a π rotation followed by a translation (i.e. an exchange), induces a +monodromy in the wave with a jump in its phase of exp(±iπ/m) (with the sign determined by the +orientation of the monodromy) in such a way the that wave function changes by a phase factor +Ψ(u, �; z1, . . ., zN) �→ e±i π +m Ψ(u, �; z1, . . . , zN) +(399) +In other words, the quasihole is an anyon with fractional statistics π/m. Daniel Arovas, J. Robert +Schrieffer and Frank Wilczek [235] further refined this argument by showing that under such +a slow adiabatic change the wave function of two quasiholes acquires a Berry phase which ac- +counts for the fractional statistics. An explicit path-integral derivation of this effect, which uses +the fact that the quasihole states are coherent states can be found in Ref. [9], chapter 13. +10.3.6. Hydrodynamic Effective Field Theory +We will now turn to the approaches to the FQHE using the methods of quantum field theory. +As we will see the concept of flux attachment plays a key role. In section 9.5 we showed that +80 + +Chern-Simons gauge theory is in fact a theory of flux attachment. Thus we expect that Chern- +Simons gauge theory should play a key role as well. +It is useful to ask first why should Chern-Simons gauge theory play a central role at least in +the description of the low energy physics. There is a simple, yet powerful, phenomenological +argument due to J¨org Fr¨ohlich and Anthony Zee that shows why this should be the case [236]. +This in essence a hydrodynamic argument. The fractional quantum Hall effect occurs in a 2DEG +in the presence of a large magnetic field which breaks explicitly time reversal invariance. In the +regime of the FQHE the 2DEG is a uniform charged incompressible fluid in which charge is +conserved. This means that the charge 3-current jµ = (j0, j) (where j0 is the charge density and +j is the charge current density) must be locally conserved, +∂µ jµ = 0 +(400) +which is the continuity equation. The solution of this equation is that the conserved current can +be written in terms of a dual (in the Hodge sense) vector field Bµsuch that +jµ = 1 +2πǫµνλ∂νBλ +(401) +The factor of +1 +2π is introduced for later convenience. The current field jµ is not changed by a +smooth redefinition of the vector field Bµ → Bµ + ∂µΦ, where Φ(x) is a non-singular field. This +means that the vector field Bµ is a gauge field. +The effective action of this theory is a functional of the current distribution µ and, hence, of +the gauge field Bµ. It should be gauge-invariant, and odd under parity and time-reversal. Since +the fluid is incompressible and uniform, at long distances and low energies the action must be a +local functional of the gauge field which should be at least Galilean invariant. In addition, the +coupling of the fluid to an external electromagnetic probe field Aµ must be of the usual form +−ejµAµ. To lowest orders in derivatives, there the unique local and gauge-invariant Lagrangian +density which is odd under time reversal (and parity) is the Chern-Simons theory +Leff[Bµ] = m +4πǫµνλBµ∂νBλ − 1 +4g2 F 2 +µν − e +2πAµǫµνλ∂νBλ + . . . +(402) +The first term is the Chern-Simons term of the Lagrangian. The coefficient m is a dimensionless +integer which, as we saw in section 9.3 is required for the theory to be invariant on a closed +manifold. The second is a Maxwell term. Here Fµν = ∂µBν −∂νAµ is the field strength tensor of +the gauge field Bµ. By power counting we see that the coupling constant g2 has units of length−1 +and, hence this term is irrelevant at long distances and will be neglected. The third term is the +coupling of the current jµ to the electromagnetic field Aµ. Upon an integration by parts this term +can be written as BµJ µ where +Jµ = − e +2πǫµνλ∂νAλ +(403) +is a current minimally coupled to Bµ. +The equation of motion of the (dynamical) gauge field Bµ is +δL +δB = 0 +(404) +which yields the relation +m +2πǫµνλBµ∂νBλ = Jµ +(405) +From the definition of the current Jµ, Eq.(403), we see that the solution of the equation of +motion of Eq.(405), up to a gauge transformation, is +mBµ = −eAµ +(406) +By plugging this relation back into the Lagrangian for the gauge field Bµ, Eq (402) (which +si equivalent to integrate out the gauge field Bµ) we find that the effective Lagrangian of the +electromagnetic field Aµ is just a Chern-Simons term +Leff[Aµ] = +e2 +4πmǫµνλAµ∂νAλ + . . . +(407) +81 + +This result allows us to read-off the Hall conductivity of the fluid +σxy = 1 +m +e2 +2πℏ +(408) +where we restored units such that ℏ is not unity. In other words, this fluid exhibits the fractional +quantum Hall effect for a fluid at filling fraction ν = 1/m. +We should note that this heuristic argument applies equally to systems of fermions, for which +m is odd, as well as to systems of bosons, for which m is even. +10.3.7. Composite Boson Field Theory +We will now discuss tow field-theoretic approaches to the FQHE. These approaches use the +concept of flux attachment as a mapping of fermions to bosons and as a mapping of fermions +to fermions. These approaches are a form of duality transformation which engineers a form +of statistical transmutation. We will first show that Chern-Simons theory can be used to effect +such a mapping. That this is possible should not be surprising since, as we saw in section 9.7, +Chen-Simons theory is a theory of fractional statistics and, hence, of anyons. +We should make some important comments on both approaches before we get into a detailed +description. While the mapping of fermions to composite bosons and fermions to composite +bosons is correct, their approximate mean field descriptions violate symmetries of the 2DEG in +a magnetic field. At the root level is the fact that the flux attachment is local in space time and +approximate descriptions of these theories bring about a large amount of Landau level mixing, +even in the large field limit. For instance, at the mean field theory level the composite boson +approach is equivalent to a Bose-Einstein condensate which is invariant under translations and +under time reversal. In this picture the breaking of time reversal is present in the coupling to a +Chern-Simons gauge field. Likewise, the mean field theory of the composite fermion theory is a +problem of composite fermions in a partially screened. While a partially screened magnetic field +still breaks time reversal, it does not behave properly under magnetic translations. As we will +see, these problems will be solved, at least in the low energy regime, by quantum fluctuations +which restore the symmetries. In both cases, in the fractional states one recovers an effective +topological field theory which captures the universal physics of these states. Needless to say it, +both approaches do poorly in the computation of non-universal dimensionful quantities such as +energy gaps, etc. These difficulties become very severe in the compressible states whose low +energy behavior is not topological. +Unlike the wave function approaches which project these states into a specific Landau level +(usually the lowest), the field theoretic approach does not effect such projection. Several papers +have been written attempting to do a field theory projected into a Landau level. These approaches +transformed the problem into quantum field theory on a non-commutative plane, which is inher- +ent the nature of the states in a magnetic field. Although some significant progress has been made +in this direction, these theories remain poorly understood [237, 238, 239, 240, 241, 242, 243] +In this section we will focus on the composite boson theory. Let us consider a theory in 2+1 +dimensions with two dynamical abelian gauge fields Aµ and Bµ, whose Lagrangian is a sum of +a Chern-Simons term at level k and a BF term: +L = k +4πǫµνλA µ∂νA λ + 1 +2πǫµνλA µ∂νBλ +(409) +The equation of motion for the field Aµ is +k +2πǫµνλ∂νA λ + 1 +2πǫµνλ∂νBλ = 0 +(410) +whose solution (up to a gauge transformation) is kAµ = −Bµ. Plugging this relation into the +Lagrangian of Eq.(409) we find that the effective Lagrangian for the field Bµ is +L[Bµ] = − 1 +4πkǫµνλBµ∂νBλ +(411) +which states that exchanging Aµ ↔ Bµ is equivalent to the “duality” k ↔ −1/k. +In section 9.7 we showed that particles coupled to a Chern simons gauge field at level k +have fractional statistics exp(±iπ/k). Hence, we see that particles coupled to the field Bµ have +fractional statistics exp(±iπk). In other words, for k odd this coupling amounts to exchanging +a theory of fermions to a theory of bosons coupled to the gauge field Bµ. This is a form of +82 + +bosonization. Alternatively, for k even it maps a theory of fermions to another theory of fermions +coupled to a gauge field Bµ. These observations have led to distinct, but ultimately equivalent +description of the quantum Hall effect. +We will begin the approach of mapping fermions to bosons and use it in the problem of the +FQHE. This is the Landau-Ginzburg theory (or composite boson theory) of Shoucheng Zhang, +Hans Hansson and Steven Kivelson [244] (ZHK). In this approach the problem of fermions +coupled to a magnetic field interacting with each other through a two-body potential (that we +will assume is ultra with coupling constant λ, for simplicity) becomes the same problem but for +a theory of bosons which are also coupled to a Chern-Simons gauge field, which we will denote +by Aµ. The Lagrangian density for this equivalent system of composite bosons is +LCB = φ∗(x)[iD0 + µ]φ(x) + 1 +2M |Dφ(x)|2 − λ(|φ(z)|4 + +1 +4πmǫµνλA µ∂νA λ +(412) +where x = (x0, x) are the space-time coordinates, µ is the chemical potential, M is the mass of +the fermions, and m is an arbitrary odd integer. The covariant derivative +Dµ = ∂µ + i e +ℏcAµ + iAµ +(413) +which effects the minimal coupling of the complex scalar field φ(x) to the background electro- +magnetic field Aµ and to the Chern-Simons gauge field Aµ, known in this context os known as +the statistical gauge field. +ZHK showed that the FQH state can be thought as being closely related to to a phase in which +the complex scalar field condenses, much as in the case of a superfluid even though the FQH iis +not a superfluid. To see how this works we will write +φ(x) = +� +ρ(x) exp(iω(x)) +(414) +The classical equations of motion of the theory of composite bosons, Eq.(412), are +δLCB +δφ∗(x) =0 +⇒ +(iD0 + µ)φ(x) − 1 +2M D2φ(x) − 2λ|φ(x)|2φ(x) = 0 +(415) +LCB +δA0(x) =0 +⇒ +1 +2πmǫi j∂iA j + |φ(x)|2 = 0 +(416) +LCB +δAi(x) =0 +⇒ +1 +2πmǫiαβ∂αA β + +i +2M +�φ∗(x)Diφ(x) − (Diφ(x))∗φ(x)� = 0 (417) +LCB +δµ =ρ0L2T +⇒ +� +d3x |φ(x)|2 = ρ0L2T +(418) +where ρ0 is the areal density of electrons. +A uniform solution of Eqs. (415)-(418) with constant amplitude of the composite bosons, +i.e. a uniform and static condensate ¯φ, with uniform statistical gauge field strength ¯ +B, requires +that the there should be no current in the ground state and that each term of Eq.(417) should be +zero. This condition implies that the external field is canceled by the average statistical field, +eB +ℏc + ¯ +B = 0. This condition together with Eqs.(416) and (418) imply that +ρ0 = 1 +m +eB +ℏc = 1/m +2πℓ2 +0 +(419) +and we conclude that the filling fraction is ν = 1 +m, which are the Laughlin states. In addition we +get that the chemical potential is µ = 2λρ0 and that |¯φ|2 = ρ0, with ρ0 given in Eq.(419). +However, these mean field results seeming imply that this system is a Bose-Einstein conden- +sate. To understand why this is incorrect, and to determine what it actually is, we need to go +beyond the mean field theor that we just described and evaluate the effects of quantum fluctua- +tions and write +φ(x) = (ρ0 + δρ(x))1/2 exp(iω(x)), +eAµ +ℏc + Aµ = δAµ +(420) +The partition function of this theory is a path integral over the fluctuating density fields δρ, the +Goldstone field ω and the fluctuation of the Chern-Simons gauge field δAµ. Upon integrating out +83 + +the massive density fluctuations to lowest (quadratic) order we find that the effective Lagrangian +for ω and δAµ is +Leff[ω, δAµ] = κ +2 (∂0ω − δA0)2 − ρs +2 (∂iω − δAi)2 + +1 +4πmǫµνλδA µ∂νδA λ + . . . +(421) +The first two terms of Eq.(421) is the effective Lagrangian of the Goldstone mode ω in the +Bogoliubov theory of superfluidity a compressibility κ and a superfluid stiffness ρs given by +κ = 1 +2λ, +ρs = ρ0 +M = ν +2πℏωc +(422) +However, as we see, this is not a superfluid since the phase field is “eaten” by the Chern-Simons +gauge field δAµ which now acquires a mass term. In this sort of Higgs mechanism the fluctuating +gauge field has a massive longitudinal mode and a massive transverse mode. Thus, this state doe +not have any massless modes. +To see that this theory describes the fractional quantum Hall effect we need to compute the +linear response to a weak electromagnetic perturbation δAµ(x). This can be accomplished by +considering the new Lagrangian +Leff[ω, δAµ, δAµ] = κ +2 +� +∂0ω − δA0 − e +ℏcδA0 +�2 +−ρs +2 +� +∂iω − δAi − e +ℏcδAi +�2 ++ 1 +4πmǫµνλδA µ∂νδA λ+. . . +(423) +and integrate out the phase field ω (which can be set to zero in the London/unitary gauge) and +the fluctuation of the statistical field δAµ to find that effective electromagnetic action is +S eff[δAµ] = +1 +4πm +e2 +ℏ +� +d3x ǫµνλδAµ∂νδAλ + . . . +(424) +from which we conclude that the Hall conductivity predicted by the composite boson theory is +σxy = 1 +m +e2 +h +(425) +as it should be for a FQHE at filling fraction ν = 1/m. +We conclude by looking at vortex states in the composite boson theory. A vortex state is +a time-independent solution of the equations of motion Eqs.(415)-(418) with the asymptotic +behavior (in the temporal gauge δA0 = 0) +lim +|x|→∞φ(x) = √ρ0 eiϕ(x) +(426) +lim +|x|→∞δAi(x) = ± ∂iϕ(x) = ±ǫi j +x j +|x|2 +(427) +where ϕ(x) is the azimuthal angle on the plane +ϕ(x) = tan−1 +� x2 +x1 +� +(428) +The energy of a neutral vortex (i.e. not coupled to a gauge field) is logarithmically divergent, +Evortex ≃ ρs +2 ln(R/a0) where R is the linear size of the system and a0 is a short-distance cutoff +(see section 5.1.1). However, the situation here is different since the complex scalar field φ(x) is +coupled to a dynamical Chern-Simons gauge field Aµ. Except for the Chern-Simons nature of +this gauge field, the problem we are dealing with is similar to that of a superconductor coupled +to a dynamical gauge field or to the Abelian-Higgs model in quantum field theory. In the case of +interest here we have finite energy vortex solutions which satisfy the asymptotic condition +lim +|x|→∞ +���� +� +i∂ j − δA j +� +φ(x) +���� +2 += 0 +(429) +which is obeyed by configurations that obey Eq.(427). Thus, at long distances, the circulation of +the Chern-Simons gauge field on a large closed contour Γ that contains the vortex satisfies +� +Γ +δA jdx j = ±2π +(430) +84 + +This vortex carries charge. To see this we compute the local charge density j0(x) +j0(x) = − δS eff +δA0(x) = −e δS eff +δδA0(x) = +e δS CS +δδA0(x) = +e +2πmǫi j∂iδA j(x) +(431) +and compute the total charge of the vortex on a larger region Σ whose boundary is Γ and obtain +Q = e +� +d2xj0(x) = +e +2πm +� +Σ +d2x ǫi j∂iδA j(x) = e +m +1 +2π +� +Γ +dx jδA j(x) = ± e +m +(432) +Therefore, the vortex of the composite boson theory has the same charge as the Laughlin quasi- +hole. +To determine the statistics of the vortex we go back to the effective Lagrangian written in the +form of Eq.(409), (from now on we set δA µ ≡ A µ) +Leff = κ +2(∂0ω−A0)2−ρs +2 (∂ jω−A j)2+ 1 +2πǫµνλA µ∂νBλ− e +2πǫµνλAµ∂νBλ− m +4πǫµνλBµ∂νBλ (433) +where Aµ is the external electromagnetic probe field, and where we used units with ℏ = c = 1. +The first two terms make the statistical gauge field Aµ massive. In the low energy limit the +statistical gauge field is frozen to the vortex configurations of the complex scalar field φ or, +equivalently, to the vortex singularities of its phase field ω. The vorticity current Ωµ is +Ωµ = ǫµνλ∂νA λ +(434) +The effective Lagrangian for the field Bµ is +Leff[Bµ] = − m +4πǫµνλBµ∂νBλ − e +2π Aµǫµνλ∂νBλ + ΩµBµ +(435) +which, except for the coupling to the electromagnetic field, has the same form as in Eq.(224), +and of the effective action introduced in section 10.3.6 on phenomenological grounds. +The effective topological field theory of Eq.(435) encodes al the universal data of fractional +quantum Hall fluids. Being topological this effective field theory has no energy scales and de- +scribes the physics at energies low compared to any excitation energy gap. In addition to yield- +ing the correct Hall conductivity σxy = νe2/h (with ν = 1/m) and the fractional vortex charge +Q = e/m, this theory will allow us to draw important additional results about the low energy +physics. +By using the results of section 9.7, we conclude that the vortices of the composite boson are +anyons with fractional statistics exp(±iπ/m), consistent with the conclusions of section 10.3.5. +In addition, on a 2-torus the ground state of this system is m-fold degenerate. This feature, +characteristic of systems with topological order, is that the ground state degeneracy depends on +the topology of the surface on which the 2DEG resides. The ground state on a torus degeneracy +was shown by Duncan Haldane and Edward Rezayi [245] by an explicit construction of a model +wave function for the Laughlin states on a torus. On a surface with g handles (i.e. genus g) the +degeneracy is mg. Xiao-Gang Wen and Qian Niu [246] showed that these results hold for any +FQH state which are, quite generally, topological quantum fluids. +Finally, we may ask how many distinct types of vortices does this theory have. The funda- +mental (Laughlin) vortex has charge e/m and statistics π/m. In principle we could have a vortex +with any integer topological charge (winding number) n ∈ Z. Such a vortex has charge ne/m and +statistics πn2/m. However, a vortex with topological charge n = m has charge e and statistics mπ. +This vortex is indistinguishable from a hole (a missing electron) since it has the same charge and +statistics. We conclude that a Laughlin state has m distinct vortices with n = 1, 2, . . ., m − 1. We +notice that this number is the as the number of degenerate states on a torus. In section 10.3.10 +we will see that this fact is closely related to the number of primary fields of a conformal field +theory associated with the FQH states. +10.3.8. Composite Fermion Field Theory +At the beginning of section 10.3.7 we noted that flux attachment can be used to define equiv- +alent theories: a) by attaching an odd number, m, fluxes to each electron thereby becoming a +composite boson, and b) by attaching an even number, m − 1, fluxes by which the electrons +turn into composite fermions, as in Jain’s construction [230]. In this section we discuss the most +salient features of the field theory approach to composite fermions, introduced by Ana L´opez and +85 + +me in 1991[247, 248] as a theory of all the Jain states. This approach was extended in 1993 by +Bertrand Halperin, Patrick Lee and Nicholas Read [232] to the case of the compressible states. +By following the same approach that we used in section 10.3.7, but adapted to the case in +which we map to a theory of a composite Fermi field ψ(x), we find their dynamics is described +by the effective action [247] +SCF = +� +d3x +� +ψ∗(x)[iD0 + µ]ψ(x) + +1 +2M |Dψ(x)|2 +� ++ +1 +4πn +� +d3xǫµνλA µ∂νA λ +− 1 +2 +� +d3x +� +d3x′ (|ψ(x)|2 − ρ0)V(x − x′)(|ψ(x′)|2 − ρ0) +(436) +where n = m − 1 is an even integer. Here V(x − x′) = δ(x0 − x′ +0)V(|x − x′|) is the instantaneous +electron-electron repulsive interaction, and ρ0 is the average areal neutralizing charge density. +As before, the covariant derivative is Dµ = ∂µ + ieAµ + Aµ, where Aµ is the electromagnetic field +(including the uniform magnetic field B), and Aµ is the statistical gauge field. We are using units +such that ℏ = c = 1. +We can simplify somewhat the form of the actions for the composite fermions by using the +fact that the Gauss law of Chern-Simons gauge theory states that the particle density and the +gauge flux are rigidly tied together, |ψ(x)|2 = +1 +2πnB ≡ ǫi j∂iA j (this is the flux attachment) as a +operator identity. Thus, we can write the action as +SCF = +� +d3x +� +ψ∗(x)[iD0 + µ]ψ(x) + +1 +2M |Dψ(x)|2 +� ++ +1 +4πn +� +d3x ǫµνλA µ∂νA λ +− 1 +2 +� +d3x +� +d3x′ +�B(x) +2πn − ρ0) +� +V(x − x′) +�B(x′) +2πn − ρ0 +� +(437) +Since this action is a quadratic form in the Fermi (Grassmann) fields ψ, we can integrate them +out to obtain the following effective action for the statistical gauge field Aµ +S eff[Aµ] = −itr +� +iD0 + µ + D2 +2M +� ++ +1 +4πn +� +d3x ǫµνλ (A µ − eAµ)∂ν (A λ − eAλ) + S int[Aµ − eAµ] +(438) +where Aµ is a probe external electromagnetic field (with vanishing average) and does not include +the uniform magnetic field. The interaction term is +S int[Aµ − eAµ] = −1 +2 +� +d3x +� +d3x′ +�(B(x) − eB(x)) +2πn +− ρ0) +� +V(x − x′) +�(B(x′) − eB(x′)) +2πn +− ρ0 +� +(439) +Here too B(x) is an external probe with vanishing average. +We will investigate the properties of the path integral for the statistical gauge field +Z[Aµ] = +� +DAµ exp(iS eff[Aµ, Aµ]) +(440) +using a saddle point expansion. The saddle-point condition of the effective action of Eq.(438) +δS eff +δAµ(x) = 0 +(441) +leads to the equation of motion for the field Aµ +⟨jF +µ (x)⟩ + +1 +4πnǫµνλ +� +F νλ(x) − eFνλ� += 0 +(442) +where ⟨jF +µ ⟩ is the expectation value of the fermionic current. In addition we need to impose the +condition for the particle density to be uniform and equal to the neutralizing background charge +⟨jF +0 (x)⟩ = ρ0. +We will assume that the electromagnetic field Aµ describes just a static uniform magnetic +field of strength B. Under these conditions, the ground state should also be static and uniform. +This means that the field strength of the statistical gauge field should be constant and uniform +value ⟨B⟩, and that the current ⟨jF⟩ should vanish in the ground state. On the other hand, the +Gauss Law of the Chern-Simons gauge field implies that ⟨B⟩ = −2πn ρ0, while the zero current +condition requires the the (statistical) electric field vanishes, ⟨E ⟩ = 0. Since the composite +86 + +fermion couples in the same way to the electromagnetic field Aµ and to the statistical field Aµ we +conclude that they experience an effective magnetic field +Beff = B + 1 +e⟨B⟩ = B − 2πn ρ0 +e +(443) +If the total number of electrons is N, then ρ0/e = N/L2 where L is the linear size of the system. +Let us denote by Neff +φ the total number of flux quanta of the effective magnetic field (in units of +ℏ = c = 1 in which the flux quantum is φ0 = 2π), and 2πNφ = BL2 is the total flux. Then, the +total effective flux is +2πNeff +φ = 2πNφ − 2πn N +(444) +Let ν = N/Nφ be the filling fraction and νeff = N/Neff +φ +the effective filling fraction. Eq.(444) +implies that these filling fractions are related by +1 +νeff = 1 +ν − n +(445) +Recall that n is an even integer. However, the system will be incompressible only if νeff = p ∈ Z. +In other words, the composite fermions fill an integer number p of the partially screened Landau +levels, which is Jain’s condition. For this condition to be satisfied the filling fraction ν(p, n) is +ν±(n, p) = +p +np ± 1 +(446) +which are the Jain fractions. In the special case p = 1 the Jain fractions are the Laughlin fractions, +with n = m − 1. Similarly we find that Beff is +Beff = ± +B +np ± 1 +(447) +So we see that the external field B is partially screened to the smaller value Beff which can be +parallel (screened) to B or anti-parallel (overscreened) to B. For the same reason, at this mean +field level the cyclotron frequency ωc is reduced by the same amount to ωeff +c = ωc/(np ± 1). +We will discuss now the effects of quantum fluctuations for the action of Eq.(438) about the +saddle-point solutions. We will work to the lowest order (quadratic) in the fluctuations, which +can be regarded as a semi-classical (or “RPA”) treatment of this theory. The quadratic effective +action for the fluctuations of statistical gauge field, which we will still denote by Aµ is +S (2) +eff[Aµ] =1 +2 +� +d3x +� +d3y A µ(x)ΠCF +µν (x, y) A ν(y) ++ +1 +4πn +� +d3x ǫµνλ(Aµ(x) − eAµ(x))∂ν(Aν(x) − eAν(x)) + S int(Aµ − eAµ) +(448) +Here, Πµν +F (x, y) is the polarization tensor of the composite fermions for p filled effective Landau +levels. The interaction term of the action is given in Eq.(439). +As required by gauge invariance, the composite fermion polarization tensor is transverse, +∂µΠCF +µν = 0, which fixes the tensorial structure. In momentum and frequency space the com- +posite fermion polarization tensor ΠCF +µν (q, ω) depends on three kernels ΠCF +0 (q, ω), ΠCF +1 (q, ω) and +ΠCF +2 (q, ω), where ΠCF +0 +and ΠCF +2 +contain the parity-even response of the composite fermions and +ΠCF +1 +is their parity-odd response. Each kernel is given as a series terms representing particle-hole +processes with simple poles at the excitations energies ωrs = (r − s)ωeff +c , with r > p (particles) +and s ≤ p (holes). Each term has a residue which is an integer power of q2 times a Laguerre +polynomial in q2. Details of these kernels can be found in Ref.[247] and in chapter 13 of Ref.[9]. +However, what will be important here is that, in the gapped states these kernels have a low energy +and low momentum limit which is local. This will enable us to find a local topological effective +action only for the gapped states. +After integrating out the statistical gauge field Aµ we find the effective action for the external +electromagnetic probe field Aµ, which has the standard form +S eff[Aµ] = 1 +2 +� +d3x +� +d3y Aµ(x) Πµν(x − y) Aν(y) +(449) +which encodes the full electromagnetic response of the system, not just of the composite fermions. +Once again, gauge invariance requires that the full polarization tensor be transverse, ∂µΠµν = 0. +87 + +In section 10.1.4 we showed that there is a relation between the polarization tensor Πµν and +the current-current correlation function Dµν. There we mentioned the these correlators are re- +quired (by gauge invariance) to satisfy a Ward identity known as the f-sum rule: +� ∞ +−∞ +dω +2π iω DR +00(ω, q) = ρ0 +M q2 +(450) +where the (retarded) density-density correlation function is related to the (retarded) polarization +tensor, DR +00(ω, q) = ΠR +00(ω, q). In the limit of low momentum q, at fixed frequency ω, The +expression for ΠR +00(ω, q) is +ΠR +00(ω, q) ≃ q2ΠR +0(ω, q) = −ρ0 +M +q2 +(ω + iǫ)2 − ω2c +(451) +where ωc = eB/(Mc) is the cyclotron frequency of the electrons. This expression saturates the +sum rule. This means that whichever corrections Eq.(451) may have must vanish in the low q +limit. In other words, in this limit the approximation that we made is actually exact. This is well +known result in the theory of the Fermi liquid [194]. +In addition, the result of Eq.(451) implies that there is a particle-hole (density) collective +mode which at low momentum has energy ℏωc, without corrections. This result is known as +Kohn’s theorem [249] which states that in a Galilean invariant system the 2DEG must have an +exact eigenstate at the cyclotron frequency. More intuitively, the center of mass of the 2DEG exe- +cutes a cyclotron motion regardless of whether the particles are free, interacting or not, fermions, +bosons, or anyons. This is a general result. Notice that the composite fermions do not satisfy +Kohn’s theorem as they cyclotron mode would be at the effective cyclotron frequency ωeff +c , and +that the leading quantum fluctuations have restored the magnetic symmetry. +In the limit of low energy ω → 0 and low momentum q → 0 of the kernels ΠCF +0 , ΠCF +1 +and +ΠCF +2 +become +ΠCF +0 (0, 0) = p +2π +M +Beff +≡ εCF, +ΠCF +1 (0, 0) = ± p +2π ≡ σCF +xy , +ΠCF +2 (0, 0) = − 1 +2π +p2 +M ≡ −χCF +(452) +where εCF is the “dielectric constant” of the composite fermions, χCF is their effective “per- +meability”, and σCF +xy is the (integer) composite fermion Hall conductivity. We then find that +low-energy effective action of the statistical gauge field Aµ is +S eff[Aµ] =1 +2 +� +σCF +xy + +1 +2πn +� � +d3x ǫµνλA µ(x) ∂ν A λ(x) +− e +2πn +� +d3x ǫµνλA µ(x) ∂ν Aλ(x) + e2 +4πn +� +d3x ǫµνλAµ(x) ∂ν Aλ(x) ++ +� +d3x1 +2 +� +εCF Ei +2(x) − χCF B2(x) +� +− +1 +8π2n2 +� +d3x +� +d3y (B(x) − eB(x)) V(x − y) (B(y) − eB(y)) +(453) +Notice that, except possibly for the last term, the effective action is a local. This is possible +because the mean field theory state is gapped. +In what follows we will focus only on the leading terms of the effective action and neglect the +subleading terms, the least two terms of the effective action which have the form of a Maxwell +term and an additional (possibly non-local) term. The remaining leading terms have a Chern- +Simons form and a BF form whose effective Lagrangian is +Leff[Aµ] = 1 +4π +� +±p + 1 +n +� +ǫµνλA µ(x) ∂ν A λ(x)− e +2πn ǫµνλA µ(x) ∂ν Aλ(x)+ e2 +4πn ǫµνλAµ(x) ∂ν Aλ(x) +(454) +If now now integrate ot the statistical gauge field Aµ we find the the electromagnetic field Aµ has +the effective Lagrangian +Leff[Aµ] = σxy +2 +ǫµνλAµ(x) ∂ν Aν(x) +(455) +from where we find that the Hall conductivity σxy of the 2DEG is +σxy = e2 +2πℏ +p +pn ± 1 +(456) +88 + +which is the Hall conductivity of the Jain states (in standard units). Here, as before, n is an even +integer. +We will now return to the effective Lagrangian of Eq.(454). As it stands, since the Chern- +Simons term has fractional level, it is not invariant under large gauge transformations and, as +such, cannot be defined on a closed manifold. To remedy this problem we can now write this +Lagrangian in terms of a theory with two dynamical gauge fields Aµ and Bµ as follows: +Leff[Aµ, Bµ] = p +4πǫµνλA µ(x) ∂ν A λ(x) − n +4πǫµνλBµ(x) ∂ν Bλ(x) ++ 1 +2πǫµνλA µ(x) ∂ν Bλ(x) − e +2πǫµνλBµ(x) ∂νAλ +(457) +This is a topological quantum field theory for all the Jain states [250]. This theory can be defined +on any manifold no matter its topology. It is easy to see that upon integrating out the field Bµ +we recover the effective field theory for Aµ of Eq.(454). Also, in the special case of the Laughlin +states, for which p = 1, we can integrate out the field Aµ, and arrive to the effective topological +field theory for the field Bµ of Eq.(435) that we derived in the composite boson theory in section +10.3.7. +Let us rewrite the Lagrangian of Eq.(457) in a more compact yet general form. To this end we +define a multicomponent gauge field A I +µ with I = 1, . . ., L, an L-component charge vector t, and +an L × L second rank symmetric matrix KIJ. Wen and Zee have given a general classification of +a large class of fractional quantum Hall states, said to be abelian, whose topological field theory +is defined by the Lagrangian [251, 252] +L = 1 +4πKIJǫµνλA I +µ (x) ∂ν A J +λ (x) − e +2πtIǫµνλAµ(x) ∂ν A λ +I (x) +(458) +In the case of the theory of Eq.(457) which has two components, L = 2, with A 1 +µ = Aµ and +A 2 +µ = Bµ. The two-component vector is t = (0, 1) and the matrix 2 × 2 matrix is +K = +�−p +1 +1 +n +� +(459) +Wen and Zee showed that, quite generally, a theory of the K-matrix form has a vacuum degener- +acy on a torus of |detK| and, on a surface of genus g, the degeneracy is |detK|g. They also showed +that the Hall conductivity is σxy = ν e2/h where the filling fraction is +ν = +L +� +I,J=1 +tI K−1 +IJ tJ +(460) +The properties of the quasiparticles can be determined by computing the appropriate correla- +tor. In the composite fermion theory, whose Lagrangian is given in Eq.(437), the gauge-invariant +composite fermion propagator is +GCF(x − y; γ(x, y)) = +� +ψ†(x) exp(i +� +γ(x,y) +dzµ(eAµ(z) + A µ))ψ(y) +� +(461) +which, as in any gauge theory, it is gauge-invariant but path-dependent. Here γ is an oriented +open path with endpoints that the space-time locations x and y. The composite fermion propa- +gator has information about many of the properties of the quasiparticles, including their electric +charge. To find what is the statistics of the quasiparticles we need to compute a two-particle cor- +relator (a four-point function) which is defined in terms of two open oriented paths, say γ1(x, y) +and γ2(u, w), with endpoints at the locations of the two particles, respectively. +The computation of these correlators is done using a feynam sum over trajectories with dif- +ferent weights which depend on the gauge field configurations. To do these calculations is, in +general, a complicated problem. However, in a gapped state, the quasiparticles are massive and +in the low energy regime these expressions are dominated by a classical trajectory on an oriented +path ˜γ(x, y) with the same endpoints x and y. The end result reduces to the computation of the +expectation value of a Wilson loop operator +W[Γ] = +� +exp(i +� +Γ +dzµ(eAµ(z) + A µ(z)) +� +A µ +(462) +89 + +on the closed oriented path Γ = γ � ˜γ− (where ˜γ−(y, x) has the opposite orientation as ˜γ(x, y) and +runs from y to x. The explicit dependence on the external electromagnetic field yields the value +of the charge of the quasiparticle through the Aharanov-Bohm effect. Likewise, the two-particle +correlator is reduced, in the asymptotic low energy regime, to the computation of an expectation +value of two Wilson loops in the effective Chern-Simons theory, as was done in section 9.7, +which yields the fractional statistics of the quasiparticles. The interested reader can find details +in chapter 13 of Ref.[9]. +A system described by a theory of the form of Eq.(458) has different types of quasiparticles. +In general we can assign an integer-valued quantum number ℓI ∈ Z as the charge of the quasi- +particle with respect to the gauge field A µ +I . The general coupling of a quasiparticle current has +the form Lqp = jµℓIA µ +I . Then, the charge Q[ℓ] and the statistics δ[ℓ] of a general quasiparticle +defined by the integer-valued L-component vector ℓ are +Q[ℓ] = −etIK−1 +IJ ℓJ, +δ[ℓ] +π += ℓIK−1 +IJ ℓJ + 1 +(463) +where the +1 is required to refer the statistics to fermions (not bosons) [250]. These results are +general and hold for any abelian FQH state [251]. +In the case of the Jain states, the charge vector is t = (0, 1) and the quasiparticle is labeled by +the vector ℓ = (1, 0) which yield the results for the charge Q and the statistics δ +Q = +−e +np + 1, +δ = π +� +−n +np + 1 + 1 +� +(464) +which are the quantum numbers of the quasiparticles of the Jain states. For the Laughlin state at +ν = 1/3 we obtains Q = −e/3 and δ = π/3 (as we should), which for the first Jain state at ν = 2/5 +the charge Q and the statistics δ are Q = −e/5 and δ = 3π/5. With some caveats, the predictions +for the quasiparticle charge in the FQH states with ν = 1/3 and ν = 2/5 have been confirmed in +noise measurements at a quantum point contact by R. de Picciotto and coworkers [253] and by +L. Samindayar and coworkers [254]. Fractional statistics has been measured experimentally in +both FQH states at ν = 1/3 and ν = 2/5 in (Fabry-Perot) interferometers by J. Nakamaura, S. +Liang, G. Gardner, and M. Manfra [233]. +10.3.9. The Compressible States +At large values p → ∞ the Jain sequences converge to the even-denominator fractions +ν∞(n) = 1/n. In this limit, Beff → 0 and ωeff +c +→ 0. In other words, at the filling fractions +ν∞(n) the average effective field experienced by the composite fermions is zero. This mean +field theory then would predict that this state is essentially a Fermi liquid of composite fermions +which is a compressible state. In addition, one can compute the effective charge of the composite +fermion and find that it is e/(np ±1), which also vanishes in the compressible limit. It is straight- +forward to see that the Fermi sea of the composite fermions is a disk with Fermi momentum +pF = (2ν∞(n))1/2ℏ/ℓ0 and that the Fermi energy is EF = ν∞(n)ℏ/(Mℓ2 +0). Halperin, Lee, and Read +[232] did an in-depth analysis of the consequences of this compressible state which are largely +in agreement with experiment. +In our discussion of the incompressible FQ states, both in the composite boson and in the +composite fermion approaches in sections 10.3.7 and 10.3.8 we saw that the mean field approxi- +mation violated symmetries that were restored by quantum fluctuations. This fact allowed us the +derive effective filed theories which are asymptotically exact in the low energy IR regime. What +allowed us to obtain these exact results is the existence of a gap. However, the compressible +states are gapless and in some sense should be regarded theories of a quantum critical system. +In fact, if we reexamine the field theory of composite fermions of section 10.3.8 we see +that, in the compressible limit p → ∞, the action of Eq.(437) describes a system of fermions at +finite density coupled to the fluctuations Aµ of the statistical gauge field with a Chern-Simons +term, but without an external uniform field (which has been canceled by the flux of the average +statistical gauge field). This theory has two salient features. one is that the only trace left of +the explicitly broken time-reversal symmetry is in the presence of the Chern-Simons term. This +means that while the mean field theory looks like a Fermi liquid, the consequences of the broken +time reversal invariance only enters in the quantum fluctuations. In addition, in this regime this +theory also breaks the particle-hole symmetry of a half-filled Landau level (in the high magnetic +field regime). +However, these contributions, already at one-loop order, typically contain IR divergencies. +These IR divergencies are due to singular forward scattering processes of the composite fermions +90 + +by the gauge fields. They are most severe in the computation of the fermion self energy and de- +pend on the form of the electron-electron interactions. For instance if the interaction potential +V(R) is short range, the one-loop contribution to the fermion self-energy has an imaginary part +withe the behavior Σ′′(ω) ∼ ω2/3 which is much larger than the real part Σ′(ω) as ω → 0 and the +the Fermi energy EF is approached [232]. This behavior, which is found in many metallic sys- +tems at quantum criticality [255, 256, 69] implies that the width of the quasiparticle is asymptot- +ically much larger than the energy as the Femi surface is approached. This one-loop result means +that perturbation theory fails in the IR and that actual behavior may be non-perturbative. This +problem is present even in the case of unscreened e2 +R (Coulomb) interactions where the singular +behavior is milder with Σ′′(ω) ∼ ω ln ω (known as Marginal Fermi Liquid scaling [257]). +At any rate, these IR singular behaviors imply that the quasiparticle picture is inadequate and +the that in these regimes there may be no quasiparticles. These systems are generally known as +“Strange Metals”. Many strongly interacting physical systems of interest, ranging form high Tc +superconductors to (largely conjectured) spin liquids with spinon Fermi surfaces to even quark- +gluon plasmas share these types of singular IR behaviors [257, 258]. In spite of a large body +of theoretical work, it has remained largely unsolved problem. Perturbative renormalization +group methods used to justify the Landau theory of the Fermi liquid [259, 260] generally fail +in these regimes, although in some cases long range interactions have allowed some degree of +control [261, 262]. Remarkably, the AdS/CFT Gauge/Gravity Duality [263] has brought new +insights into strange metal behaviors [264]. We will see below that relativistic dualities have +given additions insights into the physics of compressible states. +10.3.10. Fractional Quantum Hall Wave Functions and Conformal Field Theory +The Laughlin wave function and its generalizations share the remarkable feature of being +essentially universal. Aside form the dependence on the magnetic length ℓ0 in the gaussian +factors which ensure that the function is integrable at long distances (a feature inherited from +the single-particle states in the Landau level), the FQH “ideal” wave functions do not have any +length scales. In addition, in almost all cases, the ideal wave functions are the exact ground state +wave functions of some suitable ultra-local Hamiltonian projected onto the lowest Landau level. +Another feature, closely related to their universality and analyticity, is that these wave functions +look similar to correlation functions in two-dimensional critical (chiral) critical systems. We will +now see that these features are not accidental and point to a deep relation between quantum Hall +states and Conformal Field Theory (CFT). +Although the relation between the simpler case of the Laughlin state was known to several +people, such as Sergio Fubini [265], this deep connection with rational CFT was articulated in +considerable generality in a somewhat earlier paper by Gregory Moore and Nicholas Read [266]. +This approach showed its full power in the formulation of the non-abelian quantum Hall states. +Let us begin with the simpler case of the Laughlin states with ν = 1/m, where m is odd for +fermions and even for bosons. We will consider a U(1) Euclidean CFT of a compact boson (a +scalar field) φ(x) closely related to our discussion of bosonization in 1+1 dimensions in section +7.4. This theory has been extensively studied in the CFT literature [103, 102, 104] The Euclidean +action for this theory is +S = 1 +8π +� +d2x +� +∂µφ +�2 +(465) +This theory is compactified by the condition that the only allowed observables are invariant under +the discrete symmetry +φ(x) �→ φ(x) + 2πRn +(466) +where R is the compactification radius and n is an arbitrary integer. The observables that satisfy +this condition are vertex operators of the form +Vp(x) = exp +� +i p +Rφ(x) +� +(467) +where p is an integer. +The correlators of this theory are products of holomorphic (right-moving in Minkowski +spacetime) and antiholomorphic (left-moving in Minkowski spacetime) factors. In the context +of the FQH states we will be interested in a it chiral theory, whose correlators are holomorphic +(analytic) functions in complex coordinates z = x1+ix2. Formally, the field φ(z, ¯z) is decomposed +into a sum of a holomorphic field φ(z) and and antiholomorphic field φ(¯z), so that the propagators +is +⟨φ(z, ¯z) φ(�, ¯�)⟩ = − ln(z − �) − ln(¯z − ¯�) +(468) +91 + +In what follows we will focus on the chiral(holomorphic) field φ(z) whose propagator is +⟨φ(z)φ(�) = − ln(z − �) +(469) +The correlator of the chiral current operator +j(z) = i +R∂zφ(z) +(470) +is given by +⟨i∂zφ(z) i∂�φ(�)⟩ ∼ +1 +(z − �)2 +(471) +which tells us that this operator has (chiral) scaling dimension ∆ = 1. Similarly the correlator of +two chiral vertex operators +Vp(z) = exp +� +i p +Rφ(z) +� +(472) +is +⟨Vp(z) V−p(�)⟩ = +1 +(z − �)p2/R2 +(473) +whose (chiral) scaling dimension is ∆p = p2/(2R2). These expressions are very similar to the op- +erators that we used in section 5.1.1 to describe vortices in 2D superfluids (and planar magnets). +The main difference is that here the fields and their correlators are holomorphic where is section +5.1.1 the correlators are the product of a holomorphic and antiholormophic factors. +For general p and R the correlator of Eq.(473) has a branch cut from z to �. This means that +the monodromies of the operators, i.e. dragging an operator around the other, induces a change +in the phase of the correlator equal to 2πp2/R2. This behavior is reminiscent of the analytic +structure the the quasiparticle wave function in the FQH states and of the braiding operation +between anyons. +Moore and Read [266] showed that the following correlator in a U(1) compactified boson +CFT (with compactification radius R = √m) is equal to the Laughlin wave function at filling +fraction ν = 1/m, +Ψm(z1, . . . .zN) = +�  +N +� +i=1 +exp +� +i √mφ(zi) +� exp +� +−i +� +d2z′ √m ρ0 φ(z′) +� � +(474) += +� +i< j +(zi − zj)m exp +−1 +4 +N +� +i=1 +|zi|2 + +(475) +where we have used units in which the magnetic length is ℓ0 = 1. +In section 5.1.1 we saw that correlators of vertex operators are non-vanishing only if their +total charge (vorticity) is zero. The reason for this condition is that the (Euclidean) action is +invariant under arbitrary uniform shifts of the field φ. Now, in the correlator on the right hand side +of Eq.(475) we have a product of N vertex operators, each with the same charge √m. Then, under +an arbitrary shift by α of the field φ(z), the product of vertex operators change by a phase equal +to N √mα. On the other hand the other operator in the correlator, which describes a continuum +background charge, changes by a phase − √mρ0L2α, where L2 is the area. Since ρ0 is the areal +particle density, N/L2, the two changes of the phase cancel each other exactly. Thus, the operator +in the expectation value of Eq.(475) is charge neutral and has a non-vanishing expectation value. +The other important feature of the correlator of Eq.(475) is the choice of compactification +radius, R = √m, and of the vertex operator Vp(z) with p = m: +Vm(z) = exp(i √mφ(z)) +(476) +whose correlation function is +⟨Vm(z) V−m(�)⟩ = +1 +(z − �)m +(477) +Which is invariant under a 2π rotation (since m ∈ Z). However under a rotation by π (i.e. an +exchange) it changes by (−1)m. Hence, for m odd it changes sign, and the vertex operator Vm +behaves as a fermion, while for m even the vertex operator behaves as a boson. We will call the +vertex operator Vm ≡ Ve the electron operator. +92 + +In other words, the correlator of Eq.(475) is the expectation value of N vertex operators each +describing an electron (for m odd) and a neutralizing background charge. Then, an elementary +calculation yields the expression of the Laughlin wave function. +In this formulation the ground state wave function for the integer QH state at ν = 1, which +has the for of a Vandermonde determinant shown in Eq.(377), is interpreted as the correlator of +N electron operators in the U(1) CFT with compactification radius R = 1. In this simple case, +the electron operator is the vertex operator V1(z) = exp(iφ(z)), and there are no other operators +(aside from the current j(z) = i∂zφ). The propagator of this vertex operator is +⟨V1(z)V−1(�)⟩ = +1 +z − � +(478) +which indeed represents a fermion with electric charge e. +We will now see that the vertex operator of Eq.(472) with p = 1 (and R = √m) +V1(z) = exp +� i√mφ(z) +� +(479) +represents the Laughlin quasihole. The correlator of this vertex operator is +⟨V1(z) V−1(�)⟩ = +1 +(z − �)1/m +(480) +which has a branch cut stretching from z to � and, as a result, under a rotation by ±π its phase +changes by ±π/m, just as the Laughlin anyons do. To see that this operator indeed is related to +the Laughlin quasihole we will compute the effect of inserting the vertex operator V1(z) in the +correlator of Eq.(475) and find +� +exp +� i√mφ(�) +� +N +� +i=1 +exp +� +i √mφ(zi) +� exp +� +−i +� +d2z′ √m ρ0 φ(z′) +� � += += +N +� +i=1 +(zi − �) +� +i< j +(zi − zj)m exp +−1 +4 +N +� +i=1 +|zi|2 − 1 +4m|�|2 + += Ψm(�; z1, . . ., zN) +(481) +which is, indeed, the wave function for the Laughlin quasihole of Eq.(392). The insertion of +an additional vertex operator V1(u) at u inside the expectation value of Eq.(481) yields the two- +quasihole wave function of Eq.(397), proposed by Halperin [234] and discussed in section 10.3.5, +with the same branch cut shown in Eq.(398) and, therefore carry fractional statistics π/m. +The operator product expansion discussed in section 3.3.2 of the vertex operators of this CFT +yields new insight on the quasiholes. Indeed the OPE of two vertex operators of charges p and q +is +lim +�→u Vp(�)Vq(u) = lim +�→u +1 +(� − u)∆p+∆q−∆[p+q]m V[p+q]m(u) +(482) +where [p]m is the integer p modulo a multiple of m, i.e. if p = mr + s (with r a non-negative +integer and 0 ≤ s < m, then [p]m = s. Eq.(482) is understood in the sense that the contribution +of all additional operators to the right hand side vanish as � → u. In other words, the vertex +operators are primary fields of this CFT and there are only m primary fields. The physical process +described by the OPE is called fusion. +A CFT with a finite number of primaries is called a rational CFT [103, 102]. This result is, +of course, the same statement that we made is section 10.3.7 that the FQH state has m distinct +anyons (vortices). This result also means that a vertex operator of charge m, i.e. an electron, is +indistinguishable from the identity operator, V0. In this sense, the allowed primaries are fields +which are local respect to the electron operator Vm. From the point of view of the FQH state, an +operator that creates or destroys an electron (at fixed filling fraction) has no effect, as the FQH +state is a fluid made of electrons This also means that all physical operators must braid trivially +with the electron operator. In a CFT an operator of this type is said to generate and extended +symmetry algebra [267, 268]. +Furthermore, the OPE of the chiral current j(z) with the vertex operator V1(z) is +lim +�→u j(�)V1(u) = 1/m +z − wV1(u) +(483) +93 + +which implies that the vertex operator represents a state with charge 1/m (in units of the electric +charge e). +The CFT approach to construct FQH states has become a very powerful tool. We will see in +section 10.3.11 that FQH states on an open geometry (e.g. a disk) have edge states which can +be understood a chiral CFTs. In addition to providing a deeper understanding of more general +abelian (muliti-component) FQH states with a K-matrix structure, new classes of of FQH states +with non-abelian (multi-dimensional) representations of the Braid Group has been proposed. +The anyons of these states are of particular interest since they have been proposed originally +by Kitaev in 1997 [169] that anyons can be used as physical qubits for topologically-protected +quantum computation [169, 269, 270, 271, 170]. +So far we discussed the case of abelian anyons. Abelian anyons have the property that the +fusion of two anyons is another anyon. Similarly, a braiding operation between two anyons is +equivalent (up to a phase factor) to another anyon. Mathematically this means that abelian anyons +are one-dimensional representations of the Braid group. The fractional statistics of an anyon is +then a label of the representation of the Braid group. As we saw, the same holds under fusion. +However, the Braid group generally admits multi-dimensional representations. In this more +general case the fusion (or braiding) of two anyons is represented by a linear combination (su- +perposition) of anyon states. Linear combinations of anyon states are represented by unitary +matrices of rank greater than 1. Since matrices generally do not commute braiding and/or fusion +processes correspond to a multiplication of these matrices. Such unitary processes are then re- +garded as quantum gates. This concept is the physical basis of topological quantum computation. +It is currently the cutting edge of research in the field. +We will now discuss the simplest non-abelian FQH states, the Moore-Read (MR) states [266]. +The Moore-Read wavefunctions are +ΨMR(z1, . . . , zN) = Pf +� +1 +zi − zj +� � +i< j +(zi − zj)n exp(− 1 +4ℓ2 +0 +N +� +i=1 +|zi|2) +(484) +which describes a FQH of electrons at filling fraction ν = 1/n, with n an even integer. Here +Pf +� +1/(zi − zj +� +is the Pfaffian of the matrix 1/(zi − zj). +This state was motivated by the discovery of an unexpected FQH state in the N = 1 Landau +level with an even denominator filling fraction ν = 5/2 by Willett and coworkers in 1987 [272, +273, 274]. Until then (and since then) this is the only FQH state with an even denominator +filling fraction. The Pfaffian factor has a simple pole when two coordinates approach each other. +However, provided n ≥ 1 the “Laughlin factor” in the MR wavefunction cancels the singularity. +We will shortly discuss the CFT construction of the Moore-Read state. The poles in the MR +wavefunction suggest that in this state the electrons can get closer to each other than in a Laughlin +state. This observation suggests that there may be some form of attraction (or suppression of +repulsion) between the electrons in the MR state and motivated the notion that the observed +even-denominator plateau may be physically related to some sort of pairing interaction. Shortly +after the Moore-Read proposal was made, a paired wave function at ν = 1/2 with a Pfaffian factor +was also proposed by Greiter, Wen and Wilczek [275, 276, 277] who suggested that it reflects +electrons pairing in the ℓ = 1 angular momentum channel in time-reversal breaking condensate +with symmetry px + ipy. +Although, following the logic behind the Kohn-Luttinger mechanism for paring in angular +momentum state ℓgeq1 with repulsive interactions [278, 279, 280] it may possible to get a paired +state even in the lowest Landau level (although there is no evidence for it, so far), the N = 1 +Landau level may be more hospitable to such a state. Indeed, in the N = 1 Landau level the +single-particle states i have angular momentum greater or equal than 1, and their probability +distributions are suppressed both at short and long distances; these states look like smoke-rings. +The matrix elements of the Coulomb potential in the N = 1 Landau level are parametrically +suppressed at short distances (compared with the states in the lowest, N = 0, Landau level). In +this scenario, the ν = 5/2 plateau is interpreted as a ν = 1/2 fully polarize state of the N = 1 +landau level, with an N = 0 Landau level filled with electrons with both spin polarizations in +a state with ν = 2. There is numerical evidence that an instability into a p-wave paired state +such as the MR state is favored if the short-distance repulsion between the composite fermions is +softened [281]. +In fact, in 1983 Halperin [282] considered generalizations of the Laughlin state in which +due to very strong attractive interactions electrons would form clusters which then condense into +a Laughlin-type state. The simplest example was Laughlin state of bosons (paired electrons) +94 + +a paired state at ν = 1/2 whose Laughlin quasihole carries charge e/4. Extensive numerical +calculations are more compatible with the MR state than with the Halperin state of pairs [283]. +We will see shortly that the Halperin state is an abelian relative of the MR state. +Moore and Read derived the wave function of Eq.(484) by considering a CFT with two +sectors: a chiral boson CFT with compactification radius R = √n, and a chiral Majorana fermion +CFT, i.e. the chiral part of the 2D Ising CFT, discussed in section 4.2. The chiral Ising CFT has +central charge c = 1/2 and several primary fields: the identity I, the twist field σ, and the +(chiral) Majorana fermion χ with (chiral) scaling dimensions 0, 1/16, and 1/2, respectively. The +propagator of the Majorana fermion, which is a free field, is +⟨χ(z) χ(�)⟩ = +1 +z − � +(485) +The primary field σ, the “twist field”, is non-local with respect to the Majorana fermion χ, and +changes its boundary conditions from periodic to antiperiodic. +The fusion rules of the chiral Ising CFT are +χ ⋆ χ = I, +σ ⋆ σ = I + χ, +σ ⋆ χ = χ +(486) +What will be important below is that the σ field has two fusion channels. As in the analysis of +the Laughlin states, the compactified boson n primaries (anyons), the vertex operators Vp(x) = +exp(ipφ(x)/n), with n = 0, 1, . . .n and there are n types of anyons, and a (charge) current J = +i√n∂zφ(z). +The first task is to identify the electron operator which has to be a fermion and has to +carry electric charge. This means that it is a composite of an operators (with Fermi statis- +tics) the (neutral) chiral ising CFT and a vertex operator whose charge is an integer multi- +ple of e. +The desired electron operator is ψe(z) = χ(z) exp(i √nφ(z)) whose correlator is +⟨χ(z)Vn(z)χ(�)V−n(z)⟩ = 1/(z − �)1+n, which is a fermion with charge e. +The n-point function of the chiral Majorana fermions is +⟨χ(z1) . . .χ(zN)⟩ = Pf +� +1 +zi − zj +� +(487) +which follows fro applying Wick’s theorem. Then we see that the MR wave function is the +expectation value +ΨMR(z1, . . ., zN) = ⟨χ(z1) . . .χ(zN)⟩ × +�  +N +� +i=1 +exp +� +i √nφ(zi) +� exp +� +−i +� +d2z′ √n ρ0 φ(z′) +� � +(488) +Next we need to identify the primary fields of the tensor product of the chiral Ising CFT, Z2, +and the chiral U(1) CFT with compactification radius R = √n. The allowed primary field must +belocal with respect to the electron operator, ψe. This leaves us with four primary fields: a) the +identity I, b) the non-abelian vortex σ(z) × exp +� +i +2 √nφ(z) +� +with charge e/(2n) and non-abelian +statistics, c) the charge-neutral Majorana fermion χ, and d) the abelian vortex (the Laughlin +quasihole) exp +� +i√nφ(z) +� +, with charge e/n and abelian statics δ = π/n. +The he new feature here is the non-abelian fractional statistics of the non-abelian vortex +(sometimes called a “half-vortex”. Its non-abelian character is a consequence of the fact that the +fusions of two twist fields has two channels, see Eq.(486). This implies that the wave function of +four non-abelian vortices can be expressed as a linear combination of two so-called conformal +blocks. In other words, this wave function belongs to a degenerate two-dimensional Hilbert +space, each component labeled by a conformal block [102]. A braiding operation between two +non-abelian vortices is a monodromy of the wave function that induces a unitary transformation +U in this Hilbert space +U = +1√ +2 +exp +� +iπ +�1 +8 + 1 +4n +�� +× +� 1, +1 +−1 +1 +� +(489) +Therefore, the degenerate Hilbert space of four quasiholes provides a two-dimensional represen- +tation of the Braid Group. This observation is the basis for regarding a state of four non-abelian +quasiholes as a qubit. Furthermore, Nayak and Wilczek showed that a state of 2p quasiholes +spans a 2p−1-dimensional degenerate Hilbert space [284]. This means that, for large p, the de- +generacy per quasihole is +√ +2. In other words, this degeneracy is not due to a local degree of +freedom attached to each quasihole but that it is shared in a non-local fashion! +95 + +In what sense are the Moore-Read states paired? In an insightful paper Read and Green [285] +used a BCS theory approach to investigate the pairing properties of a condensate of composite +fermions in the px + ipy channel. The mean-field BCS Hamiltonian is +HF = +� +d2k +� +(ε(k) − µ) ψ†(k)ψ(k) + 1 +2 +� +∆∗(k)ψ(−k)ψ(k) + ∆(k)ψ†(k)ψ†(−k) +�� +(490) +where ψ(k) is the composite fermions field, ε(k) = +k2 +2M, and ∆(k) is the gap function. For a +px + ipy condensate the gap function the gap function transforms under rotations as an ℓ = −1 +angular momentum eigenstate, and for k → 0 it behaves as ∆(k) = (kx − iky)∆, where ∆ is a +constant pairing amplitude. +The BCS ground state has the form +|G⟩ = +� +k +� +u(k) + �(k)ψ†(k)ψ†(−k) +� +|0⟩ +(491) +where |u(k)|2 = |�(k)|2 = 1. The amplitudes u(k) and �(k)| satisfy the Bogoliubov-de Gennes +Equation (BdG) +� ξ(k) +−∆∗(k) +−∆(k) +−ξ∗(k) +� �u(k) +�(k) +� +≡ E(k) ˆn(k) · σ +�u(k) +�(k) +� += E(k) +�u(k) +�(k) +� +(492) +where ξ(k) = ε(k)−µ, σ = (σx, σy, σz) are the Pauli matrices, and ˆn(k) = (−Re∆(k), Im∆(k), ξ(k)) +is a unit vector defined for every k. The eigenvalues E(k) and the eigenvectors (u(k), �(k) are +E(k) = +� +ξ2(k) + |∆(k)|2, +�(k) +u(k) = −E(k) − ξ(k) +∆∗(k) +(493) +The spinor amplitudes are |u(k)|2 = 1 +2 +� +1 + ξ(k) +E(k) +� +and �(k) = 1 +2 +� +1 − ξ(k) +E(k) +� +. +In the low momentum limit and in real space the BdG Equations become +i∂tu = − µu + ∆∗i(∂x + i∂y)� +i∂t� =µ� + ∆i(∂x − i∂y)u +(494) +which is just the Dirac Equation in 2+1 dimensions, with µ playing the ole of the mass, and with +the restriction that the spinor (u, �) obeys the Majorana condition +(u, �) +�0 +1 +1 +0 +� += +�u +� +� +(495) +The Majorana condition is obeyed in all superconductors and reflects that fact that in these con- +densates only the fermion parity (−1)NF (with NF being the number of fermions) is conserved, +while the fermion number NF is not. +This is a good BCS state in that it is fully gapped and it is chiral. Assuming that the pair +fields in the px and py channels are equal, they showed that the BCS ground state |G⟩ has the +form +|G⟩ ∼ exp + +1 +2 +� +k +g(k)ψ†(k)ψ†(−k) + |0⟩ +(496) +Projected onto a state with N fermions with real space coordinates xi the wave function is a +Pfaffian +Ψ(x1, . . ., xN) = ⟨x1, . . ., xN|G⟩ = Pf(g(xi − x j)) +(497) +The long distance behavior of this wave function depends on whether the chemical potential +µ > 0 (this is the weak-pairing or BCS regime), or µ < 0 (which is the strong pairing BEC- +like regime). While in the case of an s-wave superconductor these two regimes are smoothly +connected by a crossover, in the px + ipy case they are separated by a quantum phase transition +at µ = 0. In the weak pairing regime the function g(k) has the asymptotic long distance form, as +k → 0, +g(k) ≃ − +2µ +(kx + iky)∆∗ +(498) +where the pair field behaves as ∆(k) ≃ (kx − iky)∆. In real space (in complex coordinates) the +function g(z) becomes +g(z) = +� iµ +π∆∗ +� 1 +z +(499) +96 + +which is the form used in the Pfaffian wave function. On the other hand, in the strong pairing +regime, µ < 0, g(k) has the asymptotic behavior +g(k) ≃ −A(kx − iky) +a−2 +0 + k2 +(500) +where A and a0 are functions of ∆ and µ, +A = +2|µ|M∆ +2|µ| + m|∆|2 , +a0 = +1 +2|µ| +� +2|µ| +M + |∆|2 +(501) +In real space, at distances |x| ≫ a0 the Fourier transform of g(k) decays exponentially, and as +a0 → ∞, µ → 0 and the Fourier transform g(z) has the power-law behavior +g(z) ≃ +� i|∆| +2π∆∗ +� +1 +z|z| +(502) +This means that µ = 0 is a quantum critical point that separates the weak pairing phase from the +strong pairing phase, which have distinct properties. +Read and Green then used a topological argument, originally formulated by Grisha Volovik +[286] in the context of superfluid 3He films. To see this we notice that the three component +unit vector ˆn(k) takes values on the surface of a sphere S 2. In addition, since �(k) → 0 for +k → ∞(and, hence u(k) → 1), we can wrap the momentum space k onto a sphere S 2. Thus +ˆn(k) are smooth functions of S 2 �→ S 2. such maps are classified into the homotopy classes of +π2(S 2) ≃ Z given in terms of the integer-valued topological invariant +Q = 1 +4π +� +d2k ˆn(k) · ∂kx ˆn(k) × ∂ky ˆn(k) +(503) +In the strong pairing phase µ < 0 and ξ(k) > 0 and ˆn(k) takes values only on the northern +hemisphere of the target space S 2. Such maps can be deformed continuously to the North Pole +and, hence, are topologically trivial, Q = 0. On the other hand, in the weak pairing phase, µ > 0, +ξ(k) takes both positive and negative values. In this phase ˆn(k) is topologically non-trivial and +the topological invariant Q takes non-trivial values ±1. +The upshot of this analysis is that, in the weak pairing phase, the paired FQH state is, at this +mean field level, a two-dimensional topological superconductor. Read and Green [285] further +showed that this pairing state has vortices with an interesting fermionic spectrum. A vortex is a +state win which at long distances the (complex) amplitude of the px + ipy condensate behaves as +∆ exp(iϕ), where ϕ is the azimuthal angle, which winds by 2π on a circumference of large radius +R. The BdG equation, Eq.(494), has zero mode solutions. In polar coordinates (r, ϕ) the BdG +Equation the zero modes satisfy +∆ieiϕ � +∂r + i +r∂ϕ +� +� =µu +∆ie−iϕ � +∂r − i +r∂ϕ +� +u = − µ� +(504) +The normalizable zero mode spinor solution is +�u(r, ϕ) +�(r, ϕ) +� += f(r) +√r + +1√ +ieiϕ/2 +1 +√ +−ie−iϕ/2 + +(505) +where f(r) is given by +f(r) ∼ exp +� +− +� r +0 +dr′ µ(r′) +|∆| +� +∼ exp(−µr/|∆|) +(506) +Under a 2π rotation this spinor solution is double-valued +(u(r, ϕ + 2π), �(r, ϕ + 2π)) = −(u(r, ϕ), �(r, ϕ)) +(507) +which follows form the global phase invariance. I fact, in all paired states, topological or not, un- +der a global transformation of the pair field by a phase θ, the (composite) fermion must transform +with a phase of θ/2, +∆(x) �→ eiθ∆(x), +ψ(x) �→ eiθ/2ψ(x), +ψ†(x) �→ e−iθ/2ψ†(x) +(508) +97 + +In other words, in a paired state the fermion is non-local to the vortex. Hence, under a 2π phase +transformation of the pair field, the fermions must change sign, and the spinor zero mode solution +must be double-valued. This means that this state has a brach cut. +The structure of the vortex and of the fermion zero mode is closely related to the problem of +solitons with fractional charge that we discussed in section 8.3. In fact, Roman Jackiw and Paolo +Rossi [287] investigated a closely related problem of a theory of Dirac field in 2+1 dimensions +coupled to a charged scalar field through a Yukawa coupling with the form of a Majorana mass. +They showed that such a theory admits vortex solutions with fermion zero modes. In a subsequent +paper Erik Weinberg [288] showed that these zero modes are counted by an index theorem which +relates the number of zero modes to the vorticity. More recently, Nishida, Santos and Chamon +[289] that the relativistic theory of Jackiw and Rossi in the non-relativistic approximation reduces +to the theory of a px + ipy superconductor. +There is, however, a subtle but profound difference between the fermion zero modes of the +one-dimensional fractionally charged solitons and the fermion zero modes of a superconductor. +In the case of teh 1D solitons, the zero modes can be either occupied or empty rendering a +soliton charge of −e/2 or +e/2. However, in the case of the superconductor, the BdG fermions +are charge-neutral since their charge has gone into the condensate. Thus, in a superconductor +fermion number is not conserved, and only the fermion parity is conserved. This means that +the field operator associated with the vortex zero mode, which we will denote by γ, must be a +self-adjoint operator, γ† = γ. +Ivanov [290] showed that this behavior is the origin of the non-abelian fractional statistics in +this system. Indeed, if one considers a configuration with 2n vortices, each will carry a Majorana +zero mode γi. Since these are fermion operators they satisfy the usual anticommutator algebra, +{γi, γ j} = 2δi j. We can (arbitrarily) group the 2n Majorana fermion operators into n pairs. For +each pair we can define a complex Dirac operator ψ j = (γ2 j + iγ2 j+1)/2 (and its adjoint), which +satisfies the usual fermionic algebra, {ψ j, ψ† +k} = δ jk. Each Dirac fermion can be either in an +empty state or in an occupied state. Thus a system of 2n vortices supports a degenerate Hilbert +space of dimension 2n−1, which agrees with the results of Nayak and Wilczek [284]. However, +the assignment of Majorana operators to specific pairs is actually arbitrary, which amounts to +a particular definition of the branch cuts. Changing the assignment of the operators into pairs +is then equivalent to a rearrangement of the cuts. In 2006 Michael Stone and Suk Bum Chung +[291] showed that the these vortices obey the braiding and fusion properties of Ising anyons. +These properties follow from the branch cut configurations which affect the monodromies of the +vortices. It is important to stress that the vortices with zero modes are the non-abelions. Majorana +fermions are fermions and, as such, are abelian. +In addition to the Moore-read Pfaffian state, the CFT approach has led to the formulation of +new non-trivial FQH states. Read and Rezayi [292] proposed a series of so-called cluster states, +which generalize the concept of paired states. They investigated a particular type of cluster states +in which the Pfaffian factor of the MR state is replaced by a correlator of parafermion primary +fields of a Zk CFT of Zamolodchikov and Fateev [293]. Parafermions were originally introduced +by Fradkin and Kadanoff [152] (see also Ref.[294]) as a generalization of the fermions of the 2D +Ising model to the Zk clock model. In this model one can define several types of parafermions +each consisting of the fusion of a charge operator and a magnetic (disorder) operator. These +operators obey an algebra of the type AB = exp(ip2π/k)BA, where the integer p depends on the +electric and magnetic charges of the parafermions, which is reminiscent of fractional statistics. In +close analogy with the Majorana fermions of the 2D Ising model, Zk CFT describes the behavior +of these systems their self-dual points. +The Zk CFT is actually much richer and has more primary fields than the Z2 case of the +Ising model. As a CFT, the Zk theory is the same as the CFT on the coset SU(2)−k/U(1). A +coset means that a U(1) sector has been projected out of the spectrum. Here we will focus on a +special case of the Z3 CFT. This CFT has a parafermion primary field, which we will call ψ, and +a non-abelian primary that we will call τ. Read and Rezayi [292] proposed to replace the Pfaffian +factor of the MR state with a correlator of N parafermions of a Z3 CFT and a Laughlin factor +with an exponent of n + 2/3. Such states require that the number of electrons N be a multiple of +3. Thus, these states can be viewed as sates in which the electrons cluster in groups of 3. They +also considered the more general case of the Zk CFT in which case N is a multiple of k. The +resulting filling fractions are ν = k/(mk + 2) (with m ≥ 0). The Read-Rezayi k = 3 parafermion +state is a leading candidate for the observed FQH plateau at ν = 12/5 = 2 + 2/5 [295, 296] (the +compering state being the 2/5 Jain state). There is strong numerical support for the 12/5 state +98 + +being a Z3 parafermion [297]. +The quasiholes obtained by inserting the primary fields τ in the correlator of parafermions +have interesting properties which stem from their basic fusion rule +τ ⋆ τ = I + τ +(509) +Read and Rezayi [292] showed that the number of conformal blocks for 3n quasiholes, i.e. the +number of degenerate states, in the parafermion theory is the Fibonacci number F3n−2, where +F1 = 1, F2 = 2, F3 = 3, F4 = 5. In general Fm = Fm−1 + Fm−2. For m → ∞, the rate of +increase approaches the limit limm→∞ Fm/Fm−1 = (1 + +√ +5)/2 which is known as the Golden +Mean. In other words, for large m, the dimension of the Hilbert space increases exponentially +as ((1 + +√ +5)/2)m. Aside from these interesting mathematical curiosities, the significance of this +fusion rule is that these states, regarded as“qubits”, define unitary transformations that cover +uniformly the Bloch sphere which is required for universal quantum computation [270]. +At the level of topological quantum field theory the Moore-Read and Read-Rezayi states are +related to the non-abelian Chern-Simons gauge theory with gauge group SU(2) at Chern-Simons +level k. In the more general case of the gauge group SU(N) on a manifold M action is given by +S CS[Aµ] = k +4π +� +M +d3x ǫµνλ +� +Aµ +a∂νAλ +a + 2 +3 fabcAµ +aAν +bAλ +c +� +(510) +where k ∈ Z is the level, the gauge fields Aµ = Aµ +ata take values in the algebra of SU(N), with {ta} +being the N2 − 1 generators, and fabc the structure constants of SU(N). The expectation values +of Wilson loop operators of this theory compute the Jones polynomials which are topological +invariants of knot theory [165]. The connection with SU2)k Chern-Simons gauge theory has +been essential in formulating effective low energy quantum field theories for the non-abelian +states [298, 299, 300, 301, 302, 303]. +10.3.11. Edge States and Chiral Conformal Field Theory +We will now discuss the edge states of the FQH states. As we saw the FQH states are +incompressible in the bulk and all bulk excitations are gapped. The edge of the region occupied +by the fluid (in many cases that edge of the physical sample) is where the bulk gap collapses and +hence where the system has low energy excitations. +The role of edges and their nature can be seen already in the simple case of the integer Hall +effect treated as a system of free fermions in the lowest Landau level. In this state all single +particle states are occupied, and the bulk ground state wave function is a Slater state which takes +the form of the Vandermonde determinant of Eq.(377). The potential that keeps the electrons +inside the sample increases monotonously near the edge. As we discussed in section 10.3.1, in +this region the electrons experience an electric field E that pushes them into the sample, and in +the presence of the perpendicular magnetic field B the electrons move along the edge at the drift +velocity � = c|E|/|B|. At some spatial location, corresponding to the locus of a single particle +Landau state, the potential crosses the Fermi energy and in that region potential is essentially a +linear function of the coordinate and hence of momentum (or angular momentum depending on +the gauge that is being used). In other words, the low energy states are one branch of a one- +dimensional chiral excitation, such as in the right-moving states of our discussion of the chiral +anomaly in section 7.2 +In section 9.5 we showed that a Chern-Simons gauge theory on a manifold with a boundary +projects onto a chiral CFT on that boundary. In that section we showed that an abelian Chern- +Simons theory U(1)m integrates to the boundary, the 1+1-dimensional Minkowski spacetime +S 1 × R, as a chiral U(1)m CFT for a compactified boson with compactification radius R = 1 +whose action is given by +S [ϕ] = +� +S 1×R +d2x 1 +4π +� +∂0φ∂1φ − �(∂1φ2) +� +(511) +where we rescaled the compactified scalar by a factor of √m to that its compactification radius +R = √m as in section 10.3.10. The only difference between the theory of Eq.(241) and what we +did in section 10.3.10 is that this theory is in a 1+1-dimensional Minkowski spacetime S 1 × R +while before we were in Euclidean spacetime.In the case of the integer quantum Hall effect ν = 1 +(and hence m = 1) and the (time-ordered) correlator of the vertex operator, V1(x, t) = exp(iφ(x, t)) +is +⟨T exp(iφ(x, t)) exp(−iφ(0, 0))⟩ = +1 +x − �t − iε +(512) +99 + +which is, indeed, the propagator of a chiral free fermion. +On the other hand, for the Laughlin states at filling fraction ν = 1/m (and compactification +radius R = √m) the propagator of the electron ψe ∼ exp(i √mφ) is +⟨T exp(i √mφ(x, t)) exp(−i √mφ(0, 0))⟩ ∼ +1 +(x − �t − iε)m +(513) +while the propagator of the quasihole exp +� +i√mφ +� +is +� +T exp +� i√mφ(x, t) +� +exp +� +− i√mφ(0, 0) +�� +∼ +1 +(x − �t − iε)1/m +(514) +Therefore, the CFT of the edge is the same as the CFT of the ideal wavefunction with the only +difference that the former is in Minkowski spacetime while the latter is in Euclidean spacetime. +In this sense there is a one-to-one correspondence between the bilk and the edge. We can +also see how that works by considering a fundamental Wilson arc in the bulk along a path γ(x, y) +where x and y are at the boundary +⟨W[γ]⟩ = ⟨exp(i +� +γ(x,y) +dxµA µ) ≡ ⟨exp( i√mφ(x)) exp(− i√mφ(y))⟩ +(515) +where on the right hand side we have rescaled the field by √m, as before. The scaling dimensions +of the electron, the quasihole and the current are ∆e = m +2 , ∆qh = +1 +2m and ∆current = 1. These results +will be important shortly. +The same structure applies to the multi-component abelian FQH states. The only difference is +that in multi-component FQH states the edge consists, in general, of a charge field which couples +to the electromagnetic field and one or more neutral edge states. An example is the theory of the +non-abelian Pfaffian states at filling fraction ν = 1/n. In this case the edge theory consists of a +compactified chiral boson φ of radius R = √n and a chiral Majorana (neutral) fermion χ. At a +formal level the Lagrangian for the edge state(s) is a sum of two terms +L = 1 +4π(∂xφ∂tφ − �c(∂xφ)2) + χi(∂t − �n∂x)χ +(516) +where �c and �n are the velocities of the (charged) compactified chiral scalar φ and of the neu- +tral Majorana field χ. Numerically (and experimentally) it is known that the charge mode is +(substantially) faster than the neutral mode(s), �c > �n, and often by significant factors. +A superficial reading of this Lagrangian suggests that these degrees of freedom are decou- +pled. However this is not correct. Only operators from both sectors which are local with respect +to the electron ψe ∼ χ exp(i √nφ) are physically allowed. Here an operator is local with respect +to the electron means that the operator braids trivially with the electron. This condition restricts +the allowed observables. In the case of the fermionic MR state, with n = 2, the allowed operators +are the non-abelion σ exp(iφ/2 +√ +2) with scaling dimension is ∆ = 1/8 and electric charge is +Q = 1/4, the Majorana fermion with scaling dimension ∆ = 1/2 and no electric charge Q = 0, +and the Laughlin quasiparticle with scaling dimension ∆ = 1/4 and charge Q = 1/2. Its electron +operator has scaling dimension ∆ = 3/2 and charge Q = 1. +Conformal field theory has an important defining universal quantity called the central charge +which is closely related to the energy-momentum tensor of the theory [102]. For pedagogical +introductions to CFT see Refs.[304, 45] and chapter 21 of Ref.[10]. The energy-momentum +tensor Tµν is a locally conserved current, ∂µTµν = 0. Its local conservation implies the global +conservation of energy and momentum. As such, the energy-momentum tensor is a fundamental +observable of any quantum field theory. In a conformal field theory, the energy-momentum is +also the generator of local scale and conformal transformations. In the special case of 1+1- +dimensional systems, such as the edge states of the FQH fluids, the energy momentum tensor +has special and crucially important properties. In addition of being locally conserved, conformal +invariance requires that Tµν must be traceless, T µ +ν = 0, since its trace is the generator of dilations. +In this case, if the theory is chiral, the energy momentum tensor has only one (right-moving) +component T = (T00 − T11)/2. In complex coordinates of the Euclidean metric, the correlator of +the energy momentum tensor T = Tzz is [305] +⟨T(z)T(�)⟩ = +c/2 +(z − �)4 +(517) +100 + +where c is a universal quantity known as the it central charge of the CFT. In the case of the +compactified boson the central charge is c = 1 whereas for a massless Majorana fermion c = 1/2. +The central charge of the CFT enters in many observables of fundamental physical impor- +tance. For example, for a 1+1-dimensional system of length L, such as the edge state of a FQH +droplet, the ground state energy Egnd for large L has the behavior [306, 307] +Egnd = ε0L − c π� +6L + O(1/L2) +(518) +where ε0 is the ground state energy density, which is a non-universal quantity, c is the central +charge of the CFT and � is the velocity of the massless modes. The second term in Eq.(518) +is known as the Casimir energy. Similarly, the free energy density f(T) of a CFT has the low +temperature Stephan-Boltzmann behavior (in one dimension) +f(T) = ε0 + cπT 2 +6� + O(T 3) +(519) +where we set the Boltzmann constant to unity. The low-temperature specific heat c(T) is +c(T) = cπT +3� + O(T 2) +(520) +In a chiral CFT, such as the edge states of the FQH fluids, the central charge enters in the low +temperature behavior of the (Hall) thermal conductivity κxy, [285] +κxy = cπT +6ℏ +(521) +Much of what is known about FQH states comes form experiments involving their edge +states. In a series of exquisite experiments Granger, Eisenstein and Reno [308] measured κxy in +the edge states of a ν = 1 quantum Hall fluid and showed that the heat transport is indeed chiral, +i.e. the edge state behaves as a heat “diode”. This effect was confirmed in the FQH states by +Bid and coworkers [309] who, in addition, where used this effect to detect the neutral modes in +several Jain states at filling fractions 2/3 and 2/5 and in the non-abelian state at ν = 5/2. +The simplest experimental probes are quantum point contacts and typically are of two types:inter- +edge tunneling inside a FQH liquid or tunneling of electrons into the edge state of a FQH liquid. +Since the bulk is gapped, only the edge states participate in these tunneling processes. In the +general case we will have two edges and a tunnel process at a point contact. Here the two edges +can be either the two edges of the same FQH state, in which case the process happens inside the +FQH liquid and involves tunneling of quasiparticles, or the edges of different liquids, in which +case this process is external and involves tunneling of electrons. Problems of these types were +first investigated by Charles Kane and Matthew Fisher [310] which led to a considerable amount +of work on these problems. +Let us consider first the case of tunneling into the edge state of a ν = 1/m Laughlin state +from a Fermi liquid, which we will take to be a QH fluid in the ν = 1 state. The point contact is +at x = 0 The total Lagrangian is +L = Ledge + LFL − Γ eiω0t ψ† +e,edge(0, t) ψe,FL(0, t) + h.c. +(522) +where ω0 = eV/ℏ, and V is the bias voltage between the two fluids, and Γ is a tunneling matrix +element. The local electron spectral density (the density of states) N(ω) at energy ω is +N(ω) = Im lim +x→0+ +� ∞ +−∞ +dt Ge(x, t) eiωt = const. |ω|m−1 +(523) +where Ge(x, t) = ⟨Tψ† +e(x, t)ψe(0, 0)⟩ is the electron correlator in the FQH edge, shown in Eq.(512). +This result, combined with Fermi’s Golden Rule for a point contact with voltage bias V, predicts +a tunneling current from a Fermi liquid (FL) into the chiral Luttinger liquid (CLL) of the edge +state to be [311] +I(V) = 2π e +ℏ|Γ|2 +� 0 +−eV +dE NCLL(E, T)NFL(E + eV, T) ∝ Vm +(524) +and a tunneling differential conductance G(V) +G(V) = dI +dV = 2π e +ℏ|Γ|2NFL(0)NCLL(V, T) ∝ Vm−1 +(525) +101 + +which is non-Ohmic and vanishes as V → 0. +Early experiments by Albert Chang and cowerkers , on a geometry that (most likely) had +many point contacts, confirmed the predicted CLL behavior [312, 313], although there were +discrepancies in the measured exponents. The behavior seen in more recent experiments by +Cohen and coworkers [314], on a point contact in graphene, are consistent with the theoretical +predictions [311]. Interestingly, these newer class of experiments [315, 314] also show evidence +of an analog of Andreev reflection predicted to exist near the strong coupling fixed point of the +theory of Sandler, Chamon and Fradkin [316] as a consequence of electron fractionalization. +Most experiments are done on a geometry call a Hall bar in which the QH fluid occupies a +rectangular region with its length L larger than its width W. In this geometry, there are chiral +edge states at opposite sides of the Hall bar propagating in opposite directions. One type of point +contact consists in creating a constriction in the quantum Hall fluid by applying a gate, a bias +potential on a narrow strip accross the Hall bar. This gate repels the electrons in the QH fluid, +forcing the opposite edges to approach each other in the proximity of the gate. In the absence +of the gate, momentum is conserved on each edge which forbids tunneling accross the Hall bar +since the edge states have opposite Fermi momentum. However, the gate breaks translation in- +variance on both edges and tunneling between them is now allowed. In other words, the gate +creates a point contact between the opposite propagating edge states leading to a tunneling pro- +cess between the edges of the QH liquid. Thus, this is internal tunneling as oppose the process +of tunneling between two different liquids that we discussed above. +The tunneling Lagrangian for this system has the same form as in Eq.(522) except that now +the two edges are identical. The theory Kane and Fisher shows that in the factional case the most +relevant process is the tunneling of FQH quasiparticles. The Fermi Golden Rule argument used +in Eq.(525) to compute the differential conductance also applies in the present case except that +now the two densities of states are the densities of states of the (Laughlin) quasiparticles (instead +of electrons). The local quasiparticle density of states is Nqp(ω) ∼ |ω| +2 +m −1. Then, the differential +tunneling conductance G(V) in this case is +G(v) ∼ V2( 1 +m −1) +(526) +This behavior is also non-Ohmic but, unlike the case of electron tunneling of Eq.(522), the differ- +ential tunneling conductance now diverges as V → 0. This behavior means that the Γ → 0 fixed +point is unstable and that the point contact flow to a strong coupling fixed point at Γ → ∞. Early +experiments by Milliken, Umbach and Webb in 1996 [317] showed indications of CLL behavior. +The predicted behavior was confirmed in the experiments of Roddaro and coworkers [318, 319]. +In the RG language, we can define a dimensionless tunneling amplitude g by Γ = a∆−1g, +where ∆ is the scaling dimension of the tunneling operator and a is a UV cutoff. The “tree-level” +beta function is readily found to be +β(g) = a∂g +∂a = (1 − ∆) g + O(g2) +(527) +which shows that if the scaling dimension of the operator of the tunneling particle is ∆ < 1, then +this tunneling process is relevant, while is ∆ > 1 it is irrelevant. +Kane and Fisher argued that there should be a crossover from the IR unstable weak-coupling +fixed point governed by quasiparticle tunneling to an IR stable strong coupling fixed point gov- +erned by electron tunneling. This crossover is reminiscent of the the crossover in quantum impu- +rity problems such as the Kondo problem of a magnetic impurity in a metal [320, 32, 52, 53, 321]. +The main difference is that the impurity coupling is marginal and the crossover scale, the Kondo +temperature TK, which is related to the coupling constant g as TK ∼ exp(−1/g), while in the FQH +constriction the crossover scale is TK(g) ∼ g1/(1−∆). Kane and Fisher argued that this crossover +can be viewed as a process that interpolated between a FQH fluid with a weak constriction to a +regime in which the Hall bar splits into two pieces with weak electron tunneling between them. +It turns out that, after some manipulations, the model of the constriction can be mapped into +a one-dimensional compactified boson ϕ on a semi-infinite line, x ≥ 0, with a vertex operator +acting at the boundary. The action of this system, known as boundary sine-Gordon is +S = 1 +8π +� ∞ +−∞ +dt +� ∞ +0 +dx (∂µϕ)2 + Γqp +� ∞ +−∞ +dt cos +� � +ν +2ϕ(0, t) +� +(528) +This theory has two fixed points: a) the IR unstable fixed point at Γqp = 0 where the field +Neumann boundary conditions at x = 0, ∂xϕ = 0, and b) an IR stable fixed point at Γqp → ∞ +102 + +where the field has Dirichlet boundary conditions, ϕ = 2π √2/ν n (with n ∈ Z). It turns out that +this is an integrable field theory. Paul fendley, Andreas Ludwig and Hubert Saleur [322] used the +thermodynamic Bethe Ansatz to calculate the differential tunneling conductance for the Laughlin +state at ν = 1/3 as a function of voltage V and temperature T and obtained the full weak to strong +crossover RG flow. Detailed predictions of this theory have been verified experimentally by the +work of Roddaro and coworkers [318, 319]. Point contact experiments were performed in the +more challenging ν = 5/2 FQH state by Miller and coworkers [323] and Radu and coworkers +[324]. They found that the MR state is the the one that best fits their experimental results. +The charge of particle can, in principle, be found directly by measuring the noise of a weak +current. This process is known as shot noise. In the case of the constriction, the quasiparticle +current operator is Iqp = 2eνΓqp sin +� √ν/2φ + ω∗ +0t +� +, where ω∗ +0 = eνV/ℏ. The noise spectrum, +S (ω), of the tunneling current qp is [325] +S (ω) = +� ∞ +−∞ +dt ⟨{Iqp(t), Iqp(0)}⟩ eiωt +(529) +Using the expression of the quasiparticle correlator of Eq.(514), one readily finds that, to leading +order in Γqp, the noise spectrum is +S (ω) = eν⟨Iqp⟩ + +� +1 − ω +ω∗ +0 +�2ν−1 ++ +� +1 + ω +ω∗ +0 +�2ν−1 +(530) +In the limit of zero frequency the noise spectrum takes the shot noise form +lim +ω→0 S (ω) = 2e∗⟨Iqp⟩ +(531) +where the expectation value of the tunneling current is +⟨Iqp⟩ = +2π +Γ(2ν)eν|Γqp|2 ω∗ +0 +2ν−1 +(532) +Chamon, Freed and Wen [326] calculated the exact noise spectrum for the case of a ν = 1/2 +bosonic FQH state and were able to investigate the crossover as well, and Fendley, Ludwig and +Saleur [327] used the Bethe Ansatz to construct a soliton basis to compute the DC shot noise. +These theoretical predictions were tested in tour-de-force experiments by de Picciotto and +coworkers [253] and Samindayar and coworkers [254] who were able to measure the fractional +charge of e/3 in the ν = 1/3 state and, with some caveats, e/5 in the ν = 2/5 state. Dolev and +coworkers [328] went on to measure the charge from noise experiments in the nonabelian state +at ν = 5/2 and obtained results consistent with e∗ = e/4 as predicted by theory. +We will now consider experiments on Hall bars with two quantum point contacts created by +two narrow gates transversal to the bar and separated at some distance d each other. Quantum +devices of this type are Fabry-P´erot quantum interferometers. The way they operate is as follows +[329, 298]. A current is injected in the bottom edge. At the first quantum point contact (QPC) +part if the current I1 tunnels to the opposite edge and the other part I2 goes on and tunnels at the +second QPC. The FQH fluid is assumed to occupy uniformly the region between the two QPC’s. +There is some magnetic flux Φ is that region which also contains a number of localized vortices +(quasiparticles). When both currents rejoin at the top edge they interfere and the interference has +information on the charge of the particles that tunnels through the Aharonov-Bohm effect with +the flux Φ. The interference also has information on the fractional statistics of the quasiparticles +that tunneled through their braiding with the static vortices. Chamon, Freed, Kivelson, Sondhi +and Wen proposed a setup of this type to measure the fractional statistics of the quasiparticles for +the abelian states [329]. They showed that the total current It = I1 + I2 is given by +It = e∗|Γeff|2 2π +Γ(2ν) |ω0|2ν−1 sign(ω∗ +0) +(533) +where Γeff is give by +|Γeff|2 = |Γ1|2 + |Γ2|2 + (Γ1Γ∗ +2 + Γ∗ +1Γ2) Fν +�ω∗ +0d +� +� +(534) +Here Γ1 and Γ2 are the tunneling amplitudes at the two QPCs, ν = 1/m is the filling fraction, � is +the velocity of the edge modes, d is the distance between the two QPCs and +Fν(x) = √πΓ(2ν) +Γ(ν) +Jν−1/2(x) +(2x)ν−1/2 +(535) +103 + +where Γ(z) is the Euler Gamma function and Jν−1/2(z) is the Bessel function of the first kind. In +the presence of Nq localized quasiparticles in the area between the two QPCs, the contribution +of the phases of the tunneling matrix elements get shifted to +Γ∗ +1Γ2 = ¯Γ∗ +1 ¯Γ2 exp +� +−2πi +� +ν Φ +φ0 +− νNq +�� +(536) +where φ0 is the flux quantum. In Eq.(536) the first term in the phase shift is the Aharonov- +Bohm effect of the tunneling quasiparticles, while the second term in the phase shift 2πνNq is +the contribution of the fractional statistics of the tunneling quasiparticle as its worldline braids +with the Nq localized quasiparticles. This means that there is an interference contribution to +the tunneling current (and also to the transmitted current) which is sensitive to both the charge +and to the fractional statistics of the quasiparticles. This is this effect that is being measured in +experiments. +Early attempts at doing this experiment were made by Camino and coworkers [330] but were +difficult to interpret partly due to subtle reasons related to the difficulty in controlling the actual +area of the region comprised between the two QPCs [331]. Technical advances in the fabrication +of these devices led in 2020 to the first successful measurement of the fractional statistics of the +quasiholes of the Laughlin state at ν = 1/3 (and also at the Jain state at ν = 2/5) by Nakamura, +Liang, Gardner and Manfra [233]. +The nonabelian case is more subtle both theoretically and experimentally. The theory of the +abelian interferometer of Chamon and coworkers was generalized to the nonabelian FQH states +by Fradkin, Nayak, Tsvelik and Wilczek in 1998 [298]. The structure of the interferometer is the +same as in the abelian case but the interference effects are different. In addition, in the case of +the MR state, taside from the nonabelian quasiparticle σ ∼ χ exp(iφ/2 +√ +2) has charge e/4 and +nonabelian fractional statistics, the MR state has two more abelian anyons, the charge neutral +Majorana fermion χ and the charge e/2 Laughlin quasiparticle. The number of anyons and their +properties are very different in different nonabelian FQH states. +Ignoring for now the existence of more quasiparticles, let us focus now on the fundamen- +tal anyon which is nonabelian. Fradkin and coworkers showed [298] considered a nonabelian +quasihole that is injected to the bottom edge, tunnels at the first QPC and arrives at the left end +of the top edge in state |ψ⟩. If a second such quasiparticle is now injected but now tunnels to +the top edge at the second QPC., arriving at the left end of the top edge in the state eiαBNq|ψ⟩, +where α is the Aharonov-Bohm phase determined by the flux Φ piercing the interferometer and +BNq si the braiding operator os the second quasiparticle that is circling around the Nq localized +quasiparticles in the interferometer. Then the tunneling conductance measured at the left exit of +the top edge has an interference contribution +σxx ∝ |Γ1|2 + |Γ2|2 + Re +� +Γ∗ +1Γ2eiα⟨ψ|BNq|ψ⟩ +� +(537) +The matrix element ⟨ψ|BNq|ψ⟩ is given in terms of the expectation value of the Wilson loop oper- +ators of the tunneling quasiparticles braided with the Wilson loops of the Nq localized quasiparti- +cles in the area of the interferometer. In 1989 Edward Witten [165] showed that this expectation +value, which is computed in the nonabelian Chern-Simons gauge theory, is equal to a topolog- +ical invariant of the braid known as the Jones polynomial VNq(e1π/4). Therefore, the oscillatory +component of the tunneling current (and of the conductance) measures a topological invariant! +Ref.[298] gave a general algorithm for the computation of this matrix element. However, in +the particular case of the MR states, the non-abelian gauge theory associated with these states is +the SU(2)2 Chern-Simons gauge theory. In this case, an explicit calculation Bonderson, Kitaev +and Shtengel [332] and by Stern and Halperin [333] leads to the result +σxx ∝|Γ1|2 + |Γ2|2, +for Nq odd +σxx ∝|Γ1|2 + |Γ2|2 + 2 |Γ1||Γ2|(−1)Nψ cos +� +α + arg +�Γ2 +Γ1 +� ++ π +4 Nq +� +, +for Nq even +(538) +Here, Nψ = 1 when the Nq quasiparticles fuse into the state ψ and Nψ = 0 otherwise. The +interference effect is absent for Nq odd since an odd number of σ quasiparticles cannot fuse into +the identity state I. This simple even-odd effect is special for SU(1)2. In the general case and, +in particular in the SU2)3 case which applies to the k = 3 Read-Rezayi (“Fibonacci”) state, the +expressions are more complex. +104 + +At any rate, even in the MR state the situation is more complicated for two reasons. One +is that the charge e/2 abelian Laughlin quasiparticle is always able to tunnel thus spoiling the +even-odd effect. the other complication is that the nonabelian FQH occurs in the N = 1 Landau +level and the edge states of the nonabelian FQH state is surrounded by an abelian ν = 2 state, +which makes accessing the interesting edge states more difficult. Robert Willett has pioneered the +fabrication and operation of the interferometer for the nonabelian state at ν = 5/2 [334, 335, 336]. +With some significant caveats although there is solid evidence of nonabelian braiding, more work +remains to be done on this system. +11. Particle-Vortex Dualities in 2+1 dimensions +11.1. Electromagnetic Duality +The oldest form of duality in Physics is, perhaps, Dirac’s observation that in the absence of +electric charges and currents Maxwell’s equations are invariant under the exchange of electric +and magnetic fields, E → B and B → −E [114]. This observation led him to conjecture the +existence of magnetic monopoles. In a relativistic invariant formulation, Maxwell’s equations in +free space can be expressed in terms of the electromagnetic field tensor Fµν and the dual field +tensor F∗ +µν +∂µFµν = 0, +∂µF∗ +µν = 0 +(539) +where F∗ +µν = 1 +2ǫµνλρFλρ, and where ǫµνλρ is the fourth-rank antisymmetric Levi-Civita tensor, The +first Maxwell equation in Eq.(539) is just the wave equation in free Minkowski spacetime. The +second equation in Eq.(539) is known as the Bianchi identity. The Bianchi identity is a constraint +which implies that there are no magnetic monopoles and that the second rank antisymmetric field +strength tensor can be expressed in terms of the vector potential Aµ as Fµν = ∂µAν − ∂νAµ. +This is an example of what in differential geometry is called Hodge duality, which relates a +vector or, more generally a tensor field to its Hodge dual. In general in D dimensions the Hodge +dual of a (fully antisymmetric) tensor of rank p is an antisymmetric tensor of rank D − p. When +contracted with the oriented infinitesimal element of a p-dimensional hypersurface, dx1 ∧ dx2 ∧ +. . . ∧ dxp, an antisymmetric tensor of rank p, Fµ1µ2...µp, defines a p-differential form, called a p- +form. Thus, differential forms embody the physically intuitive notions of circulation of a vector +field, flux of a second rank tensor, etc. We will see below that duality transformations are closely +related to these geometric notions of duality. +11.2. Particle-Vortex Duality in 2+1 dimensions +In this section we will extend the particle-vortex duality discussed in 2D in section 5.1.1 (see +Eq.(75)) to 3D. In the field theory interpretation, 2D is a 1+1-dimensional Minkowski spacetime +and the XY model is a representation of a complex scalar field of unit modulus. In the duality, the +particles are the particle-like excitations of the complex scalar field. The high temperature phase +of the XY model is viewed as a partition function of a set of oriented loops that carry charge. the +loops are the worldlines of the particles of the XY model. +11.2.1. The 3D XY Model +This picture can be seen as follows. Consider an XY model on a D-dimensional hypercubic +lattice whose sites are labelled by {r}. At each site there is a periodic variable θ(r) ∈ [0, 2π). The +partition function is +ZXY = +� +r +� 2π +0 +dθ(r) +2π +exp + +1 +T +� +r +� +j=1,...,D +� +cos ∆jθ(r) +� +(540) +where ∆jθ(r) = θ(r+e j)−θ(r), with j = 1, . . ., D, is the lattice difference, and T is the temperature +(in the classical statistical mechanical picture). Since the interaction on each link is a periodic +function of the phase difference ∆jθ(r), we can expand the Gibbs weight (for each link!) in a +Fourier series or, what is the same, in the integer-valued representations of the group U(1) (which +is the global symmetry of the XY model) to obtain +ZXY = +� +r +� 2π +0 +dθ(r) +2π +∞ +� +ℓj(r)=−∞ +exp +− +� +r, j +T +2 ℓ2 +j(r) + i +� +r +θ(r)∆jℓ j(r) + +(541) +105 + +where we used a Gaussian approximation for the modified Bessel function In(z), and ∆jℓ j(r) = +�D +j=1(ℓ j(r)−ℓ j(r−e j)) (where e j is the lattice unit vector along the direction j) denotes the lattice +divergence. Integrating-out the phase variables θ(r) we obtain +ZXY = +� +r +∞ +� +ℓj(r)=−∞ +� +r +δ(∆jℓ j(r)) exp +− +� +r +D +� +j=1 +T +2 ℓ2 +j(r) + +(542) +Therefore, the partition function is given by a sum over loops of conserved currents ℓ j(r) defined +on the links of the lattice, with a weight on each link ∝ exp(−ℓ2 +j/2β). These loops are the (lattice) +worldlines of the particles of the complex scalar field. In the phase where this representation +is convergent, the complex scalar field is massive, the XY model is gapped, and these particles +have short range interactions. This picture is true in all dimensions, 3D included. +On the other hand, in 2D the vortices are point-like events in Euclidean spacetime which in- +teract with each other through long range, logarithmic, interactions. In the field theory language, +in 2D the point-like vortices are instantons which govern the low temperature phase of the XY +model. In spacetime dimensions D > 2 the vortices become extended objects: vortex loops (or +strings) in 3D, closed vortex surfaces in 4D, etc. In 3D the vortex loops are magnetic flux tubes +which interact with each other through a Biot-Savart type interaction, namely bits of vortices +interact with each other with a 3D Coulomb interaction much in the same way as with loops of +current in magnetostatics, +Z3DXY = +� +{mj(˜r} +� +˜r +δ(∆jm j(˜r)) exp +−2π2 +T +� +˜r,˜r′ +3 +� +j=1 +m j(˜r)G0(˜r − ˜r′)m j(˜r′) − α +� +˜r, j +m2 +j(˜r) + +(543) +where {˜r} labels the sites of the dual (cubic) lattice, the variables m j(˜r) take values on the integers, +α is a (short-distance) vortex core energy, and G0(˜r − ˜r′) is the 3D lattice propagator (Green +function) which at long distances has the standard form +G0(x − x′) = +1 +4π|x − x′| +(544) +In Eq.(543) the vortex loops are represented by the integer-valued conserved currents m j(˜r), +which are naturally interpreted as the world lines of magnetic charges. +We could have also reached the same result by following the line of reasoning that we used in +section 5.1.1. Indeed, let θ(x) be the phase field of a complex scalar field φ(x) deep in its broken +symmetry state where the amplitude of the field φ(x) can be taken to be approximately fixed. The +partition function in this phase reduces to +Z[aµ] = +� +Dθ exp +� +− 1 +2g +� +d3x +� +∂µθ − aµ +�2� +(545) +where g is a coupling constant which in the XY model is proportional to the temperature. Here, +much as in Eq.(70), the gauge field aµ represents the vortices. Indeed, in 3D the vorticity is +represented by a locally conserved vector field ωµ +ωµ = ǫµνλ∂νaλ = 2π +� +k +mk +µδ(x − xk) +(546) +where mk +µ are the integer-valued vortex (magnetic) currents we used above. As before, the pe- +riodic nature of the phase field θ(x) implies that the vortex currents must be quantized and be +integer-valued. +Here too we can perform a Hubbard-Stratonovich transformation in terms of a vector field bµ +to find a dual theory which now is +Z(aµ) = +� +DbµDθ exp +� +−g +2 +� +d3x b2 +µ − i +� +d3x bµ(∂µθ − aµ) +� +(547) +Upon integrating out the phase field θ, which acts as a Lagrange multiplier which imposes the +constraint ∂µbµ = 0. This constraint implies that we can write the field bµ in terms of a dual +gauge field ϑµ +bµ = ǫµνλ∂νϑλ +(548) +106 + +Therefore, we can rewrite the partition function as +Z(aµ) = +� +Dϑ exp +� +−g +4 +� +d3x f 2 +µν + i +� +d3x ωµϑµ +� +(549) +where fµν = ∂µϑν − ∂νϑν is the field strength of the dual gauge field ϑµ. Notice that the vortex +current ωµ is minimally coupled to the dual gauge field ϑµ. If we now integrate-out the gauge +field ϑµ we obtain an expression for the weight in the path integral for a configuration of vortices +mk +µ which is identical to the result of Eq.(543). +What we have shown above is that in 3D the Goldstone phase of a complex scalar field +with coupling constant g is the dual of a Maxwell gauge theory. with coupling constant 1/g. +Moreover, the periodicity of the phase field implies that the dual gauge field ϑµ is compact in +the sense that its fluxes must obey flux quantization. It is straightforward to show that a charge +operator Vn = exp(inθ) with electric charge n in the scalar field theory is represented in the dual +gauge theory by a magnetic monopole of magnetic charge n. +In summary we showed that the 3D XY model (equivalent to the theory of the complex scalar +field) can be written in terms of two different models of loop configurations. We saw that the +XY model is a theory of particle (electric) loops in the high temperature phase and of vortex +(magnetic) loops in the low temperature phase. However the particle loops have short range +interactions while the vortex loops have long range interactions. Thus, the 3D XY model is not +self-dual. In spite of having a representation in terms of vortex loops, the 3D XY model is in +reality has a very different behavior that the 2D system. The Kosterlitz-Thouless theory describes +the phase transition in terms of a vortex-antivortex unbinding transition upon which the vortices +proliferate. In addition, in 2D the Goldstone phase is actually a line of fixed points, there is never +true long range order but, instead, power-law correlations of the physical observables. In the +Kosterlitz-Thouless theory the disordered phase arises the strong fluctuations of the phase field +due to the proliferation of vortices while, at the local level, the order parameter still has a finite +magnitude. A consequence of this behavior is that the superfluid density has a universal jump at +the Kosterlitz-Thouless transition [337]. +In contrast in the 3D XY model the Goldstone phase is a true phase with long range order. +The theory has a continuous phase transition (the thermodynamic superfluid transition) at which +the superfluid density vanishes continuously. This means that the phase transition of the 3D XY +is better described by the Wilson-Fisher fixed point of a complex scalar field [30, 48], rather than +by a vortex proliferation picture of the Kosterlitz-Thouless theory [84, 85]. In particular, since +this theory does not have magnetic monopoles, the vortex loops cannot proliferate as they do in +2D. Instead, as the phase transition is approached, the vortex loops grow large in size but also +become fractal-like objects whose effective core size diverges as the transition is approached. In +fact, qualitative vortex proliferation arguments readily lead to the incorrect conclusion that the +transition should be (strongly) first order [338]. +In spite of these differences, particle-vortex duality still plays an important role in 3D by +relating the 3D XY model to another theory which is its dual under electromagnetic duality. Here +electromagnetic duality is understood as the exchange of the electric worldlines (electric loops) +and the magnetic loops (vortex loops) while exchanging strong and weak coupling, T ↔ 1/T, in +this case between different theories. This duality was investigated by Michael Peskin [339], by +Paul Thomas and Michael Stone [338], and by Chandan Dasgupta and Bertrand Halperin [340]. +11.2.2. Scalar QED in 3D +These authors considered a theory of a superconductor represented by a complex scalar field +ϕ(x), coupled to a fluctuating electromagnetic (Maxwell) field, also known as the abelian Higgs +model, or scalar electrodynamics (scalar QED) aµ, with µ = 1, 2, 3 The partition function of this +theory is +ZSC = +� +DϕDϕ∗Daµ exp +� +− +� +d3x L[ϕ, ϕ∗, aµ] +� +(550) +where +L = |Dµ(a)ϕ|2 + m2|ϕ|2 + u|ϕ|4 + 1 +4e2 f 2 +µν +(551) +where the covariant derivative is Dµ(a) = ∂µ + iqaµ, with the integer q being the charge of the +scalar field (in units of the coupling constant e), and fµν = ��µaν − ∂νaµ is the field strength. This +theory is known as (Euclidean) scalar electrodynamics (or, the abelian Higgs model). +107 + +The scalar electrodynamics was extensively studied using the perturbative renormalization +group within the epsilon expansion (near 4 dimensions) which leads to the conclusion that it +has a weakly (fluctuation-induced) first order phase transition [109, 341]. Scalar QED of an N- +component scalar field coupled to a Maxwell gauge field has a continuous phase transition with +a non-trivial fixed point for N ≥ Nc ≃ 183 (!) [342]. Since these results rely on perturbation +theory (or in the large-N limit) it was long suspected that the physics may be different in D = 3. +We will see that particle-vortex duality provides the answer to this question [340]. +We begin by writing the lattice version of Eq.(550) which is obtained by coupling the XY +model of Eq.(540) to a dynamical (Euclidean) Maxwell field +ZSC = +� +r +� 2π +0 +dθ(r) +2π +� ∞ +−∞ +daµ(r) +2π +exp +� +−S (θ, aµ) +� +(552) +where the action S (θ, aµ) is +S (θ, aµ) = − 1 +T +� +r,µ +cos +� +∆µθ(r) − qaµ(r) +� ++ 1 +4e2 +� +r,µ,ν +� +∆µaν(r) − ∆νaµ(r) +�2 +(553) +In this action we assumed that the abelian gauge fields do not have monopole configurations. +The partition function of Eq.(550) admits a representation as a sum over loops. We can then +used the same line of reasoning that led to Eq.(542) and write the partition function as a sum over +loops of the worldlines of the particles (charges) of the complex scalar field +ZSC +Zgauge +0 += +� +r +∞ +� +ℓµ(r)=−∞ +δ(∆µℓµ(r)) exp +− +� +r,µ +T +2 ℓ2 +µ(r) + +� +exp +iq +� +r,µ +ℓµ(r)aµ(r) + +� +a +(554) +where ⟨O[a]⟩a is the expectation value of the operator O[a] over the gauge fields aµ, and Zgauge +0 +is +the partition function of the free gauge fields. After computing this free-field expectation value +we obtain the result +ZSC +Zgauge +0 += +∞ +� +{ℓµ(r)}=−∞ +δ(∆µℓµ(r)) exp +− +� +r,µ +T +2 ℓ2 +µ(r) − q2e2 +2 +� +r,µ +� +r′,ν +ℓµ(r) Gµν(r − r′) ℓν(r′) + +(555) +where +Gµν(r − r′) = ⟨aµ(r)aν(r′)⟩ +(556) +is the Euclidean propagator of the free gauge fields. Since the configurations of the worldlines, +the currents represented by ℓµ(r), are conserved and satisfy the local constraint ∆µℓµ = 0, the +second term in the exponent of Eq.(555) is gauge invariant. Hence, we can use the propagator in +the Feynman gauge +Gµν(r − r′) = δµνG0(r − r′) +(557) +where G0(r− r′) is the 3D lattice propagator which has the same Coulomb form at long distances +already given in Eq.(544). +Therefore we can write the partition function of the scalar QED model (the superconductor) +as a sum over worldline loop configurations of the particles of the scalar field in the equivalent +form +ZSC +Zgauge +0 += +∞ +� +{ℓµ(r)}=−∞ +δ(∆µℓµ(r)) exp +− +� +r,µ +T +2 ℓ2 +µ(r) − q2e2 +2 +� +r,r′,µ +ℓµ(r) G0(r − r′) ℓµ(r′) + +(558) +which is the same as the partition function of the 3D XY model in the broken symmetry state, +given in Eq.(543), with the identification of the particle loops of the abelian Higgs model with the +vortices of the 3D XY model and a relation between the coupling constants. Hence we obtain the +duality between the broken symmetry phase of the 3DXY model (“low T”) and the symmetric +phase of the abelian Higgs model (“high T”) +ℓµ ↔ mµ, +q2e2 ↔ 2π2 +T , +T +2 ↔ α +(559) +where the quantities on the left hand side of the identifications refer to the 3D abelian Higgs +model and the quantities on the right hand side to the 3D XY model. +108 + +We can now repeat the same analysis but in the broken symmetry of the abelian Higgs model. +In this phase the gauge field aµ becomes massive by the Higgs mechanism (or, what is the same, +by the Meissner effect of the superconductor), and in this phase the vortex loops mµ of the +complex scalar field have short range interactions. This phase then is mapped onto the unbroken +phase of the 3DXY model with the screened vortex loops of the abelian Higgs model identified +with the particle loops of the 3D XY model, and with the same identifications between the +coupling constants of the two theories given in Eq.(559). The identification between the two +partition functions implies that the phase transitions must be the same. In other words, the +duality implies that the 3D abelian Higgs model has a continuous phase transition with the same +fixed point as that of the complex scalar field in 3D. The only difference is that the phases are +reversed and, for this reason, this is sometimes called an “inverted” 3D XY transition [340]. +11.2.3. The duality mapping +We can summarize these results with the following identification of two Lagrangians (in real +time, Minkowski signature)[208] +|∂µ(B)φ|2 − m2|φ|2 − u|φ|4 ↔ |∂µ(a)ϕ|2 + m2|ϕ|2 − u|ϕ|4 − 1 +4e2 f 2 +µν + 1 +2πǫµνλaµ∂νBλ +(560) +The left hand side of this equivalence is the Lagrangian of the 3D complex scalar field φ and Bµ +is a background electromagnetic gauge field. The right hand side is the Lagrangian of the abelian +Higgs model with a scalar field ϕ and a U(1) gauge field aµ. Notice that the mass terms of the +two sides are inverted reflecting the reverse order of their two phases. This identification also +shows that the current of the scalar field jµ = − i +2(φ∗∂µφ − φ∂µφ∗) is identified as +jµ ↔ 1 +2πǫµνλ∂νaλ +(561) +The field operator of the complex scalar field creates worldlines (as opposed to loops) of charged +particles. In the dual abelian Higgs theory this operator is identified with an operator that creates +a magnetic monopole of the U(1) gauge field aµ with unit magnetic charge. It is easy to see that in +the broken symmetry phase of the abelian Higgs model a monopole-antimonopole pair creation +operator decays exponentially with distance due to the flux expulsion effect (the Meissner effect +of a superconductor). This means that the monopoles are confined with a linear potential. This +is the same behavior of the complex scalar field φ in its symmetric phase where the propagator +of the complex scalar field decays exponentially with distance. Conversely, in the broken sym- +metry phase of the complex scalar field theory, the field operator φ condenses and its propagator +approaches the value |⟨φ⟩|2 at long distances. This is the same behavior one readily finds for the +monopole operator of the abelian Higgs model in the symmetric phase . +11.3. Bosonization in 2+1 dimensions +The particle-vortex duality that we discussed in section 11.2 essentially has been understood +since the 1980s. In section 7.4 we discussed in detail the problem of bosonization in 1+1 dimen- +sions whic is a mapping between a massless Dirac fermion and a massless compactified scalar +field. As we noted, this mapping is a powerful theoretical tool to investigate non-perturbatively +the structure of many theories in 1+1 dimensions of great physical interest. This success has +motivated a sustained effort to extend these concepts to higher dimensions where the problem in +many ways is more subtle. +Fermi systems in dimensions higher differ from the 1+1-dimensional case in two significant +ways. Due to the kinematic restriction of one space dimension a massless relativistic fermion +and a theory of fermions (relativistic or not) at finite density are mostly equivalent to each other. +The reason is quite simple. Consider a free massless Dirac theory in 1+1 dimensions. As we +saw, it is equivalent to a theory of two chiral fermions, the right and left moving component of +the spinor. If we add a chemical potential µ, the right and left moving states will be filled up to +µ, which is the Fermi energy. While in space dimensions d > 1 a finite chemical potential leads +to a finite fermi surface, in d = 1 space dimensions the Fermi “surface” is just two points, for +the right and left moving fermion modes respectively. The Dirac Hamiltonian at finite chemical +potential is +H = ψ†(−i)α∂xψ − µψ†ψ +(562) +109 + +Under a chiral transformation with angle θ(x), the field operators change as ψR → eiθψR and +ψL → e−iθψL and the Hamiltonian becomes +H = ψ†(−i)α∂xψ − (µ − ∂xθ) ψ†ψ +(563) +Clearly if we choose ∂xθ = µ (or, what is the same, θ = µx) the explicit dependence on the +chemical potential has been cancelled and the transformed Dirac Hamiltonian is the same as the +Hamiltonian at zero chemical potential. However under this transformation the mass operators +transform as ¯ψψ → cos(2µx) ¯ψψ and i ¯ψγ5ψ → sin(2µx) i ¯ψγ5. In the non-relativistic context this +change amounts to a modulation of the charge density with wave vector Q = 2pF(µ). +In contrast, in space dimensions d > 1 the massless Dirac system and a system of fermions +at finite density (relativistic or not) are no longer equivalent to each other as the latter system has +a Fermi surface (of co-dimension d − 1) wile the former system does not. This difference has +profound effects on their dynamics: the system of fermions at finite density is, at least in the weak +coupling regime, is expected to be a Fermi liquid (except for an instability to a superconducting +state even for infinitesimal attractive interactions) while the massless Dirac theory is stable as all +local interactions are irrelevant. +11.3.1. Bosonization of the Dirac theory in 2+1 dimensions +We will now turn to the problem of Bose-Fermi mappings in relativistic systems in 2+1 +dimensions. This problem is of great interest both in Condensed Matter Physics and in High +Energy Physics. +Bosonization in 2+1 dimensions is a more subtle problem than in 1+1 dimensions that we +discussed in section 7.4. There we saw, in 1+1 dimensions bosonization consists of a set of +operator identities relating two free field theories, a theory of massless Dirac fermions and a +compactified massless free scalar field φ(x). The success of that program is largely based on the +fact that both theories are massless (and hence are scale and conformally invariant), on the chiral +anomaly, and on the relation between the gauge current of the Dirac theory with the topological +current of the compactified boson. +The physics in 2+1 dimensions (and in general) is very different. For instance, instead of +the chiral anomaly in 2+1 dimensions theories of relativistic fermions have a parity anomaly, +discussed in section 10.1.6. We will also see that although in 2+1 dimensions there are theories +of free massless Dirac fermions and free massless compactified bosons they are no longer dual to +each other. For these and other reasons it has been difficult to extend the bosonization program +to 2+1 dimensions. +One of the first bosonization constructions for a theory of massive (relativistic) Dirac fields +was put forth by Alexander Polyakov in 1988 [168, 167]. For reasons that no longer relevant, +Polyakov considered a theory of CP1 complex scalar fields (in its unbroken phase) coupled to +a U(1)1 Chern-Simons gauge theory. Nevertheless, his main argument, with some corrections +associated with the parity anomaly, still holds correct. +Here we will follow the work of Hart Goldman and myself [343] and consider instead a +theory of a single massive complex field coupled to a U(1)1 Chern-Simons gauge theory whose +Lagrangian density is +L = |Dµ(Aµ)φ|2 − m2|φ|2 − λ|φ|4 + 1 +4πǫµνλA µ∂νA λ +(564) +In essence, this theory is a relativistic version of what we did in section 10.3.7 where we mapped +non-relativistic fermions to (also non-relativistic) bosons whereas here the theory is relativistic +and we are mapping bosons to fermions. In his work Polyakov actually considered the problem +of the propagator of a free massive scalar field coupled to the Chern-Simons gauge field. In this +limit, the propagator G(x, x′) is the transition amplitude from x to x′ (in Minkowski spacetime). +In the case in which φ(x) is a free massive field this propagator can be expressed in the for of a +Feynman path-integral as a sum over all open oriented paths {Px,x′} with endpoints at x and x′ +which in Euclidean spacetime is +Gγ(x,x′) = +� +{Px,x′ } +exp(−mL(Px,x′)) +� +exp +� +i +� +Px,x′ +dzµA µ(z) +� � +U(1)1 +(565) +where the expectation value is over the Chern-Simons gauge fields Aµ of Eq.(564). +110 + +Let us consider now the partition function of the bosons which is a sum over all closed paths +{P}, which we will assume to be non-intersecting (which requires a short distance repulsion) +Z = +� +{P} +exp(−mL(P)) +� +exp(i +� +P +dzµA µ(z)) +� +U(1)1 +(566) +The expectation value is given by +� +exp +� +i +� +P +dzµA µ(z) +� � +U(1)1 = exp +� +−1 +2 +� +P +� +P +dxµdyν⟨Aµ(x)Aν(y) +� +≡ exp(iπW(P)) +(567) +where W(P) is the writhe of the closed path P. The computation of the writhe requires a regu- +larization. One possible regularization is to thicken the path which is equivalent to including a +Maxwell term for the gauge field in the Lagrangian. In general, the writhe of the path P is +W(P) = S L(P) − T(P) +(568) +where S L(P) is an integer-valued topological invariant called the self-linking number and T(P) +is the twist (or torsion) of the path. The twist T(P) is a Berry phase which depends on the +coordinates of the space in which the path P is embedded, and it is not a topological invariant. +We will now compute the Berry phase T(P). Let ˆe(s) be a unit vector tangent to the path +P and where we parametrized the closed path P by a coordinate s ∈ [0, L]. The closed path +P is the boundary of a surface Σ we can take to be a disk. We will write the Berry phase by +extending ˆe(s) smoothly to the interior of Σ as ˆe(s, u), where 0 ≤ u ≤ 1 and ˆe(s, u = 1) = ˆe(s) +and ˆe(s, u = 0) = ˆe0 ≡ constant. With these definition the Berry phase is +T(P) ≡ W(ˆe) = 1 +2π +� L +0 +ds +� 1 +0 +du ˆe · ∂sˆe × ∂uˆe +(569) +which is defined modulo an integer. +In this formulation, in the bosonic theory the amplitude for a path of length L with tangent +vector ˆe(s) with endpoints at x and x′ is +G(x − x′) = +� ∞ +0 +dL +� +D ˆe(s) δ +� +1 − |ˆe|2� +δ +� +x − x′ − +� L +0 +ds ˆe(s) +� +exp(−|m|L ± iπW(ˆe)) (570) +In momentum space we find +G(p) = +� ∞ +0 +dL +� +Dˆe δ(1 − |ˆe|2) exp(−|m|L ± iπ W(ˆe)) exp(ipµ +� L +0 +ds ˆeµ(s)) +(571) +By inspection of Eq.(569) we see that this is the same expression that we looked in the theory of +the path integral for spin in section 4.3.2. for spin S = 1/2 particle in a magnetic field bµ = ±2pµ. +The equation of motion for ˆe is +∂sˆeµ = ±2ǫµνλˆeλ = i[H, ˆeµ] +(572) +where H = ∓pµˆeµ is the Hamiltonian. Upon quantization, ˆeµ satisfies the commutation relations +[ˆeµ, ˆeν] = 2iǫµνλˆeλ +(573) +This means that we should make the identification ˆeµ → σµ where σµ are the three 2 × 2 Pauli +matrices. Up to rescaling of the mass m → M, upon performing the path integral of Eq.(571) we +find that G(P) is the (Euclidean) Dirac propagator [168] +G(p) = +1 +ipµσµ − M +(574) +Using these results the partition function of Eq.(566) becomes +Zfermion = det[i/∂ − M] = +� +DJ δ(∂µJµ) exp �−|m|L[J] − isgn(M)πΦ[J]� +(575) +where Φ[J] = W[J], defined in Eq.(568). +111 + +We will return to this construction shortly below when we include the effects of the parity +anomaly. +Early attempts at deriving a mapping for a theory of Dirac fermions in 2+1 dimensions were +based on their behavior in the presence of gauge fields and the associated parity anomaly, dis- +cussed in section 10.1.6. These early theories are essentially a hydrodynamic description of the +Dirac theory deep in the massive phase. With minor differences, the same results were derived +independently (and simultaneously) by two groups, Fidel Schaposnik and myself [344] and Cliff +Burgess and Francisco Quevedo [345]. This approach was reexamined more recently by A. Chan, +T. Hughes, S. Ryu and myself in 2016 in a derivation of a hydrodynamic effective field theory for +topological insulators in different dimensions [346, 347]. These derivations are similar in spirit +to what we discussed in section 10.3.6 for the fractional quantum Hall effect. We will follow the +approach of Ref.[346] for the special case of a Chern insulator 2+1 dimenxions. +The Dirac theory, in any dimension, is globally gauge invariant. By Noether’s theorem this +means that it has a locally conserved current, jµ = ¯ψγµψ which thus satisfies ∂µ jµ = 0. Here +too, this means that we can write jµ = ǫµνλ∂νbλ, where bµ is a gauge field. We expect that the +effective action of the gauge field bµ should be local, gauge invariant and, in this case, relativistic +invariant. In 2+1 dimensions the effective action should break time reversal invariance and hence +it should have a Chern-Simons term. +We will consider the free massive Dirac theory coupled to a background probe gauge field Aµ. +The Lagrangian is L = ¯ψ(i/(D) − M)ψ, where Dµ = ∂µ + iAµ. As we saw in section 10.1.6, gauge +invariance and locality require that we have an even number of Dirac fermions. Here we will +assume that one of the Dirac fermions is massive and acts as a regulator. The partition function +Z[Aµ] = +� +D ¯ψDψ exp(iS F[ ¯ψ, ψ, Aµ]) +(576) +By definition the expectation value of a product of current operators jµ can be obtained by func- +tional differentiation of the partition function with respect to Aµ. +The partition function is gauge invariant, i.e. +Z[Aµ = Z[Aµ + aµ] +(577) +where the vector field aµ is a pure gauge, aµ = ∂µφ and fµnu = ∂µa − ν − ∂µaµ = 0. Hence, aµ is +said to be flat. Therefore, up to a normalization +Z[Aµ] = +� +D[aµ]pure Z[Aµ + aµ] += +� +Daµ δ(fµν = 0) Z[Aµ + aµ] += +� +DaµDbµ Z[Aµ + aµ] exp +� +− i +2 +� +d3x ǫµνλ bµ fνλ +� +(578) +where in the last equality we introduced a representation of the delta function and the vector field +bµ plays the role of a Lagrange multiplier field. using the invariance of the integration measure +under aµ → aµ − Aµ we find +Z[Aµ] = +� +DaµDbµZ[aµ] exp +� +− i +2 +� +d3x ǫµνλ bµ (fνλ − Fνλ) +� +(579) +where Fµν is the field strength of the external field Aµ. From these identities and by differentiation +of Eq.(579) it follows that a general expectation value of products of currents is +⟨jµ(x) jν(y) . . .⟩ = +δ +δAµ(x) +δ +δAν(y) . . .Z[Aµ] = ⟨ǫµαβ∂αbβ ǫνγδ∂γbδ . . .⟩ +(580) +which means that, at the operator level, we can identify [344] +jµ(x) ⇔ ǫµνλ∂νbλ +(581) +This result is actually general. The only difference is the nature of the field b isn different di- +mensions: in 1+1 is a pseudo-scalar, in 2+1 is a vector (gauge) field, in 3+1 is an anti-symmetric +(Kalb-Ramond) tensor field, etc. +112 + +Returning to the partition function of Eq.(579) we find, using the result of Eq.(324) applied +the partition function Z[aµ], that Z[Aµ] is given by +Z[Amu] = +� +DaµDbµ exp(i +� +d3x Leff[aµ, bµ, Aµ]) +(582) +where the effective Lagrangian is +Leff = −ǫµνλ bµ ∂νaλ + s +4πǫµνλaµ∂νaλ − +1 +4geff +fµν f µν + Aµǫµνλ∂νbλ + . . . +(583) +where fµν = ∂µaν − ∂νaµ. Here s = 1 in the Chern insulator, s = 0 in the trivial insulator and +s = 1/2 at the quantum critical point, and geff an effective coupling constant (with dimensions of +length−1). Here we included the effects of the parity anomaly. We recognize that the first term of +the effective Lagrangian of Eq.(583) is a BF term. +In Ref.[346] a similar result is also derived for a 3+1-dimensional topological insulator. The +main differences are that there is no parity anomaly but an axial anomaly and that the Lagrange +multiplier field is a Kalb-Ramond field, +Leff[aµ, bµν, Aµ] = ǫµνλρ bµν ∂λaρ + +θ +8π2 ǫµνλρ∂µaν∂λaρ − 1 +4g2 fµν f µν + Aµǫµνλρ∂νbλρ + . . . (584) +where θ = π in the topological insulator and θ = 0 in the trivial insulator. +Although these results are correct, they do not give a full bosonization mapping. In particular, +these results do not identify a fixed point for the dual bosonic theory and, along with it, a full +mapping of the observables. Progress on this problem has only been achieved in the past few +years both in the high energy literature [348, 349, 350, 351, 352, 199, 353, 354], and in the +condensed matter physics literature [355, 356, 357, 358, 359]. Much of that new insight was +presented in a 2016 insightful paper by Nathaniel Seiberg, T. Senthil, Chong Wang and Edward +Witten [208]. Many of the dualities are actually conjectures supported by strong consistency +checks. A derivation of the basic bosonization duality based on loop models was constructed by +Hart Goldman and myself [343]. +The basic conjectured bosonization duality is a mapping of a free Dirac fermion to a gauge +complex scalar field with a Chern-Simons term [352, 208]. We can think of the Dirac theory as +either being defined entirely in 2+1 dimensions or as a being defined at the boundary of a non- +trivial 3+1 dimensional topological insulator. In the first scenario we need to take into account +the contribution of the fermionic doublers (this is what happens in the case of a lattice theory) +or as the Pauli-Villars heavy fermionic regulators. In both cases, the additional heavy degrees of +freedom cancel the anomaly of the 2+1-dimensional Dirac fermion, see section 10.1.6. In the +second scenario, there is only one Dirac fermion at the boundary and the bulk θ-term cancels the +anomaly of the boundary theory, see section 10.2.3. With these provisos, the Lagrangian LA of +the free massive (or massless) Dirac fermion in 2+1 dimensions coupled to the electromagnetic +gauge field Aµ is +LA = ¯ψ(i /D(Aµ) − M)ψ − 1 +8πǫµνλAµ∂νAλ +(585) +where M is the Dirac mass. We will call this Theory A. Here /Dµ(A) = γµ(∂µ − iAµ) and Aµ +is a background (non-dynamical) gauge field. The last term in Eq.(585) is the Chern-Simons +term with the 1/2-quantized coefficient, the usual short-hand for the η-invariant term of Eq.(342) +needed to cancel the parity anomaly. +From the effective field theories we discussed above we know that the fermionic current +maps to the curl of a gauge field, see Eq(581). Then, the conjectured bosonic dual is given by the +Lagrangian of Eq(564) with an extra term for the (dual) coupling to the electromagnetic gauge +filed. We will call this Theory B whose Lagrangian LB is +LB = |Dµ(Aµ)φ|2 − m2|φ|2 − λ|φ|4 + 1 +4πǫµνλA µ∂νA λ + 1 +2πǫµνλAµ∂νA λ +(586) +To check the consistence we first observe that both theories are anomaly free. By functional +differentiation of the two partition functions we check that we get the correct mapping for the +fermionic current, jµ ↔ ǫµνλ∂νA λ. This mapping implies that the charge density of Theory A +maps onto the gauge field flux of Theory B, which is electromagnetic duality. We see that this +bosonization is a relativistic version of flux attachment. +113 + +Let us now identify the mapping of the phases of both theories. If the Dirac mass M < 0, +Theory A describes an anomalous quantum Hall insulator. Integrating out the massive fermions +we fund that the effective action for the electromagnetic gauge field Aµ is a U(1)1 Chern-Simons +theory. This means that in this phase σxy = −1/(2π) (in units in which e = ℏ = c = 1). Looking +now at Theory B we see that if m2 > 0, the scalar field is in the unbroken phase, and it is massive. +In this phase we set φ = 0 and find that the low energy resulting theory if just a U(1)1 Chern- +Simons gauge theory coupled to the curl of the electromagnetic field. Upon integrating out the +Aµ gauge field we find that the effective action of Aµ is also a U(1)1 Chern-Simons term, which +implies that σxy = −1/(2π). +Conversely, for M > 0 the effective electromagnetic action of the fermionic theory is a +Maxwell term (the Chern-Simons term canceled). This phase is a trivial insulator. Looking now +at Theory B we see that for m2 < 0 this theory is in its Higgs phase where ⟨φ⟩ � 0 and the gauge +field Aµ now has a mass term, ∝ A 2 +µ . In the low energy limit Aµ → 0, and the Hall conductivity +vanishes, consistent to what is expected from Theory A. +On the other hand, for M = 0 Theory A is at a quantum critical point with a very simple CFT +structure but a non-vanishing Hall conductivity σxy = 1/(4π). It is natural to map this CFT to +a Wilson-Fisher (WF) fixed point of Theory B. At the WF one sets the (renormalized!) mass +m2 +R = 0 (not the bare mass). In the absence of the Chern-Simons gauge field this WF fixed point +in well understood form high quality (five loop!) epsilon expansion calculations (enhanced with +Borel resummation) [49], and by more recent numerical Conformal Bootstrap methods [360]. +However, not much is known of the gauged version of the CFT of Theory B since it cannot +be accessed by either methods. If the mapping to Theory A is correct it should have a Hall +conductivity of σxy = 1/(4π). This conjectured value hast not yet been confirmed. +Consider now a monopole operator of the gauge field Aµ of Theory B. We will denote this +operator MA . The Chern-Simons term is not gauge invariant in a monopole background. To put +it differently it has charge 1 under the Chern-Simons gauge field Aµ. It also has charge 1 under +the electromagnetic field Aµ. Likewise the scalar field φ has charge 1 under the Chern-Simons +gauge field Aµ. The operator φ†MA is gauge-invariant under the Chern-Simons gauge field Aµ +(i.e. it has charge 0). This composite operator has electromagnetic charge 1 (which it inherited +from the monopole). Moreover, it has spin-1/2 required by the Wu-Yang construction [151]. +In other words, the operator φ†MA has the same quantum numbers as the Dirac fermion and +are identified by this conjecture. Notice, however, that the free massless Dirac field has scaling +dimension ∆ψ = 1 in 2+1 dimensions. The conjecture implies the composite operator φ†MA +should also have scaling dimensions 1 at the (gauged) WF fixed point of Theory B. Many of these +conjectures have been verified in non-abelian versions of Theory A and Theory B: a theory of +free massless Dirac fermions coupled to a Chern-Simons gauge theory with gauge group SU(N)k +and a theory of complex scalars coupled to a non-abelian Chern-Simons gauge theory with gauge +group SU(k)N, both in the limits N → ∞ and k → ∞ with N/k fixed [348, 349, 350, 351, 352]. +The particle-vortex duality discussed in section 11.2 combined with the fermion-vortex du- +ality we sketched here provide a web of dualities. These identifications constitute a powerful +non-perturbative tool which has led to many significant developments. For instance, Goldman +and myself [361] used the web of dualities to explain the experimentally observed self-duality +at fractional quantum Hall plateau transitions, which was a long standing puzzle. It has also +provided a powerful new tool to derive effective field theories of non-abelian fractional quantum +Hall states [301] and even to propose novel states with a single (Fibonacci) anyon [303]. +11.3.2. Bosonization of the Fermi Surface +A form of bosonization has been developed for the case of systems of fermions with a Fermi +surface. Dense Fermi systems at sufficiently weak coupling are well described by the Landau +theory of the Fermi liquid [13, 3, 194, 260, 259]. In this regime, the system of fermions has a +collective mode, a bound state of particles and holes, which at long wavelengths is well described +by the random phase approximation (RPA) [362]. However, in space dimensions d > 1 there is no +kinematical restriction and quasi-particles and quasi-holes may move in their separate ways. The +result is that, in addition to the collective modes, there is a low-energy spectrum of renormalized +but essentially free quasi-particles. Because of the existence of this quasi-particle spectrum the +collective modes generally (but not always) decay into particle-hole pairs resulting in a finite +lifetime of the collective modes. +Superficially the particle-hole collective modes are similar to the scalar field of the bosonized +theory in d = 1. To an extent it has been possible to “bosonize the Fermi surface” and to treat it +as a quantum mechanical object [363, 364, 365, 366, 367, 368]. In this approach one essentially +114 + +regards each direction normal to the Fermi surface as a one-dimensionalchiral fermion, including +the current algebra structure, subject to global constraints that cancels the anomalies. This point +of view ensures that the total fermion number in the Fermi sea is conserved. Here I will only +present a short summary of the main ideas. +As in the one-dimensional case one defines a filled Fermi sea state |FS⟩. This is a state of non- +interacting fermions filling up all one-particle states up to the Fermi energy EF. For concreteness +we consider a system in two space dimensions. In this case, neglecting lattice effects, the Fermi +surface is a circumference of radius pF, the Fermi momentum. At fixed fermion number, the +excitations are particle-hole pairs. The operator nk(q) = c†(k+ q +2) c(k− q +2) creates a particle-hole +pair with relative momentum q with total momentum k. In the low energy regime k is a point on +the Fermi surface (with |k| = pF) and q is small momentum compared with pF. In what follows +we will label the point k on the Fermi surface by the angle θ of the arc spanned by k and an +arbitrary origin on the Fermi surface. +We will normal-order the particle-hole creation operator with respect to the filled Fermi sea +and define δ(q, θ) =: nk(q) := nk(q) − ⟨FS|nk(q)|FS⟩. Haldane [363], Houghton and Marston +[364, 367], and Castro Neto and Fradkin [365, 366] showed that in real space that these normal +ordered density operators obey the equal-time commutation relations +[δn(x, θ), δn(x′, θ′)] = − 1 +2π kF(θ) · ▽δ2(x − x′)δ(θ − θ′) +(587) +where kF(θ) is a unit vector normal to the Fermi surface at angle θ. This is the algebra of the +quantum fluctuations of the Fermi surface. We can further define at each point θ on the Fermi +surface a Bose field ϕ(x, θ) such that +δn(x, θ) = N(0)�F(θ) · ▽ϕ(x, θ) +(588) +where N(0) is the density of one-particle states at the Fermi surface and �F(θ) is the Fermi veloc- +ity at the location θ. The scalar field ϕ(x, θ) is a chiral boson at θ which parametrizes the quantum +fluctuations of the Fermi surface. The quantum dynamics of the chiral bosons ϕ(x, θ) is governed +by the action +S =1 +2N(0) +� 2π +0 +dθ +2π +� +d2x dt +� +−∂tϕ(x, θ) �F(θ) · ▽ϕ(x, θ) − (�F(θ) · ▽ϕ(x, θ))2� ++1 +2N(0) +� 2π +0 +dθ +2π +� 2π +0 +dθ′ +2π +� +d2x d2x′ dt F(x − x′; θ − θ′) �F(θ) · ▽ϕ(x, θ) �F(θ′) · ▽ϕ(x, θ′) +(589) +where x = (x, t), and F(x−x′; θ−θ′) are the Fermi liquid parameters that parametrize the effective +forward scattering interactions among the fermion quasiparticles on the Fermi surface [13]. The +equation of motion predicted by this (quadratic) action is equivalent to the linearized quantum +Boltzmann equation familiar from the Landau theory of the Fermi liquid. A non-linear extension +has been introduced recently by Delacr´etaz and coworkers [368]. Recent work on the role of +quantum anomalies in dense Fermi systems has yielded new insights on these problems [369]. +In the regime, where the Landau theory of the Fermi liquid is expected to work [13], bosoniza- +tion of the Fermi surface approach has reproduced the previously known results. There remain +many open problems in this approach (and others) in the vicinity of quantum phase transitions. In +spite of intense research using many different approaches, quantum phase transitions in metallic +systems are not yet fully understood beyond perturbation theory [255, 256]. Non-perturbative +ideas such as deconfined quantum criticality have been proposed [370] and significant work has +been done using large-N methods [371, 372]. A particularly important metallic quantum phase +transition (and perhaps the simplest) occurs near the Pomeranchuk instability of the fermi liquid. +A quantum phase transition to an electron nematic state [373] has been predicted and studied +within the Hertz-Millis approach. It has also been studied using higher dimensional bosoniza- +tion [373, 374, 375]. Quantum Monte Carlo simulations [376] have shown that the vicinity +of a nematic quantum critical point can trigger a superconducting state. Nematic Fermi fluids +have been found in many physical systems of interest ranging from high temperature supercon- +ductors (such as cuprates and iron superconductors), to electron gases in large magnetic fields +[377, 378, 379, 380]. +115 + +12. Conclusions +In this chapter I have attempted to cover the role of Quantum Field Theory in modern Con- +densed Matter Physics. Its role and influence is vast and deep. In spite of the length of this +chapter, I have not done justice to its role in many important areas of Condensed Matter. In +particular, I have not touched on the the theory of quantum spin liquids, fractons, and particu- +larly topological phases in three space dimensions, among others. An area in which the ideas +and concepts of QFT have a huge impact , both conceptually and as a tool, is on the problem of +quantum entanglement. The concept of quantum entanglement as applied to condensed matter +systems, both in equilibrium and out of equilibrium, is a major area of current development. It +has become a crucial tool for characterizing systems at quantum criticality as well as topological +phases of matter. Nevertheless, I hope that the reader will find this chapter both insightful and a +useful resource. +Acknowledgments +This work was supported in part by the US National Science Foundation through grant No. +DMR 1725401 at the University of Illinois. +References +[1] J. R. Schrieffer, Theory of Superconductivity, Frontiers in Physics, Addison-Wesley, New York, New York, 1964. +[2] Y. Nambu, G. Jona-Lasinio, Dynamical Model of Elementary Particles Based on an Analogy with Superconduc- +tivity. I, Physical Review 122 (1961) 345. +[3] A. A. Abrikosov, L. P. Gorkov, I. E. Dziyaloshinski, Methods of Quantum Field Theory in Statistical Physics, +Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 1963. +[4] A. L. Fetter, J. D. Walecka, Quantum Theory of Many-Particle Systems, McGraw-Hill, New York, New York, +1971. +[5] S. Doniach, E. H. Sondheimer, Green’s Functions for Solid State Physicists, Frontiers in Physics, W. A. Benjamin +Inc., New York, New York, 1974. +[6] R. P. Feynman, A. R. Hibbs, Quantum Mechanics and Path Integrals, McGraw-Hill, New York, New York, 1965. +[7] R. P. Feynman, Statistical Mechanics, A Set of Lectures, Frontiers in Physics, W. A. Benjamin Inc., Reading, +Massachusetts, 1972. +[8] L. D. Faddeev, Introduction to Functional Methods, in: Methods of Field Theory, Proceedings of the Les Houches +Summer School 1975, Session XXVIII, R. Balian and J. Zinn-Justin editors, North Holland, Amsterdam, 1976, +pp. 1–40. +[9] E. Fradkin, Field Theories of Condensed Matter Physics, Second Edition, Cambridge University Press, Cam- +bridge, U.K., 2013, (expanded 2nd edition of the originally published book in 1991). +[10] E. Fradkin, Quantum Field Theory: An Integrated Approach, Princeton University Press, Princeton, N. J., 2021. +[11] A. Altland, B. D. Simons, Condensed Matter Field Theory, 2nd Edition, Cambridge University Press, Cambridge +U.K., 2010. +[12] D. Pines, P. Nozi`eres, The Theory of Quantum Liquids, Vol. I, W. A. Benjamin Inc., New York, New York, 1966. +[13] G. Baym, C. J. Pethick, Landau Fermi Liquid Theory, John Wiley & Sons, New York, NY, 1991. +[14] L. P. Kadanoff, G. Baym, Quantum Statistical Mechanics: Green’s Function Methods in Equilibrium and Non- +Equilibrium Problems, Frontiers in Physics, W. A. Benjamin Inc., New York, New York, 1962. +[15] A. Kamenev, Field theory of Non-Equilibrium Systems, 1st Edition, Cambridge University Press, Cambridge, +U.K., 2011. +[16] P. C. Martin, Measurements and Correlation Functions, Gordon and Breach Science Publishers Ltd., New +York/London/Paris, 1968. +[17] D. C. Mattis, The Theory of Magnetism, Harper & Row, New York, New York, 1965. +[18] L. D. Landau, On the Theory of Phase Transitions, Zh. Eksp. Teor. Fiz. 7 (1937) 19. +[19] L. D. Landau, E. M. Lifshitz, Statistical Physics, Vol. 5 of Course of Theoretical Physics, Pergamon Press, Oxford, +U.K., 1959. +[20] L. Onsager, Crystal Statistics. I. A Two-Dimensional Model with an Order-Disorder Transition, Physical Review +65 (1944) 117. +[21] T. D. Schultz, D. C. Mattis., E. H. Lieb, Two-Dimensional Ising Model as a Soluble Problem of Many Fermions, +Rev. Mod. Phys. 36 (1964) 856. +[22] E. Lieb, T. Schultz, D. C. Mattis, Two soluble models of an antiferromagnetic chain, Annals of Physics (N. Y.) 16 +(1961) 407. +[23] P. Pfeuty, The one-dimensional Ising model with a transverse field, Annals of Physics 57 (1970) 79. +[24] C. N. Yang, The Spontaneous Magnetization of a Two-Dimensional Ising Model, Physical Review 85 (1952) 808. +[25] L. P. Kadanoff, Scaling Laws for Ising Models near Tc, Physics 2 (1966) 263. +[26] L. P. Kadanoff, Spin-spin correlations in the two-dimensional Ising model, Il Nuovo Cimento B 44 (1966) 276. +[27] E. Efrati, Z. Wang, A. Kolan, L. P. Kadanoff, Real-space renormalization in Statistical Mechanics, Rev. Mod. +Phys. 86 (2014) 647–667. +[28] K. G. Wilson, Renormalization Group and Critical Phenomena. I. Renormalization Group and the Kadanoff Scal- +ing Picture, Physical Review B 4 (1971) 3174–3183. +[29] K. G. Wilson, Renormalization Group and Critical Phenomena. II. Phase-Space Cell Analysis of Critical Behavior, +Physical Review B 4 (1971) 3184–3205. +[30] K. G. Wilson, M. E. Fisher, Critical Exponents in 3.99 Dimensions, Phys. Rev. Lett. 28 (1972) 240. +116 + +[31] K. G. Wilson, J. B. Kogut, The Renormalization Group and the ǫ Expansion, Physics Reports C 12 (1974) 75. +[32] K. G. Wilson, The Renormalization Group: Critical phenomena and the Kondo problem, Rev. Mod. Phys. 47 +(1975) 773. +[33] K. G. Wilson, The Renormalization Group and Critical Phenomena, Rev. Mod. Phys. 55 (1983) 583. +[34] A. Z. Patashinskii, V. L. Pokrovskii, Behavior of Ordered Systems Near the Transition Point, Soviet Physics JETP +23 (1966) 292, Zh. Ekzp. i Teor. Fiz. 50, 439 (1966). +[35] M. E. Fisher, The Theory of Equilibrium Critical Phenomena, Reports on Progress in Physics 30 (1967) 615. +[36] M. E. Fisher, The renormalization group in the theory of critical behavior, Reviews of Modern Physics 46 (11974) +597–615. +[37] L. P. Kadanoff, W. G¨otze, D. Hamblen, R. Hecht, E. A. S. Lewis, V. V. Palciauskas, M. Rayl, J. Swift, D. Aspnes, +J. Kane, Static Phenomena Near Critical Points: Theory and Experiment, Rev. Mod. Phys. 39 (1967) 395–431. +[38] M. Gell-Mann, F. E. Low, Quantum Electrodynamics at Small Distances, Physical Review 95 (1954) 1300. +[39] N. N. Bogoliubov, D. V. Shirkov, Introduction to the Theory of Quantized Fields, Interscience, New York, New +York, 1959. +[40] L. P. Kadanoff, Operator Algebra and the Determination of Critical Indices, Phys. Rev. Lett. 23 (1969) 1430. +[41] K. G. Wilson, Non-Lagrangian Models of Current Algebra, Phys. Rev. 179 (1969) 1499. +[42] A. M. Polyakov, Conformal Symmetry of Critical Fluctuations, JETP Letters 12 (1970) 381, [ZhETF Pis. red. 12, +538 (1970)]. +[43] A. M. Polyakov, Non-Hamiltonian approach to quantum field theory, Soviet Physics JETP 39 (1974) 10, [Zh. +Eksp. Teor. Fiz. 66, 23 (1974)]. +[44] L. P. Kadanoff, H. Ceva, Determination of an Operator Algebra for the Two-Dimensional Ising Model, Phys. Rev. +B 3 (1971) 3918. +[45] J. Cardy, Scaling and Renormalization in Statistical Physics, Cambridge Lecture Notes in Physics, Cambridge +University Press, Cambridge, U.K., 1996. +[46] G. ’t Hooft, M. Veltman, Regularization and Renormalization of Gauge Fields, Nuclear Physics B 44 (1972) 189. +[47] C. G. Bollini, J. J. Giambiagi, Dimensional Renormalization: The Number of Dimensions as a Regularizing +Parameter, Il Nuovo Cimento B 12 (1972) 20. +[48] D. J. Amit, Field Theory, the Renormalization Group and Critical Phenomena, McGraw Hill, New York, NY, +1980. +[49] J. Zinn-Justin, Quantum Field Theory and Critical Phenomena, 4th Edition, International Series of Monographs +in Physics, Oxford University Press, Oxford, U.K., 2002. +[50] D. J. Gross, F. Wilczek, Ultraviolet Behavior of Non-Abelian Gauge Theories, Phys. Rev. Lett. 30 (1973) 1343– +1346. +[51] P. W. Anderson, A poor man’s derivation of the scaling laws of the Kondo problem, J. Phys. C: Solid State Phys. +3 (1970) 2436. +[52] N. Andrei, Diagonalization of the Kondo Hamiltonian, Physical Review Letters 45 (1980) 379. +[53] P. B. Wiegmann, Exact solution of the Kondo problem, JETP Letters 31 (1980) 392. +[54] H. D. Politzer, Reliable Perturbative Results for Strong Interactions?, Phys. Rev. Lett. 30 (1973) 1346. +[55] K. G. Wilson, Confinement of Quarks, Physical Review D 10 (1974) 2445. +[56] J. Kogut, L. Susskind, Hamiltonian formulation of Wilson’s lattice gauge theories, Phys. Rev. D 11 (1975) 395. +[57] M. Creutz, L. Jacobs, C. Rebbi, Monte Carlo computations in lattice gauge theories, Physics Reports 95 (4) (1983) +201–282. +[58] A. M. Polyakov, Interaction of goldstone particles in two dimensions. Applications to ferromagnets and massive +Yang-Mills fields, Physics Letters B 59 (1975) 79. +[59] S. H. Shenker, J. Tobochnik, Monte Carlo renormalization group analysis of the classical Heisenberg model in +two dimensions, Phys. Rev. B 22 (1980) 4462. +[60] E. Br´ezin, J. Zinn-Justin, Renormalization of the Nonlinear σ Model in 2 + ǫ Dimensions: Application to the +Heisenberg Ferromagnets, Phys. Rev. Lett. 36 (1976) 691. +[61] D. H. Friedan, Nonlinear Models in 2 + ǫ Dimensions, Annals of Physics 163 (1985) 318. +[62] P. W. Anderson, Absence of Diffusion in Certain Random Lattices, Phys. Rev. 109 (1958) 1492. +[63] E. Abrahams, P. W. Anderson, D. C. Licciardello, T. V. Ramakrishnan, Scaling Theory of Localization: Absence +of Quantum Diffusion in Two Dimensions, Phys. Rev. Lett. 42 (1979) 673. +[64] F. J. Wegner, The mobility edge problem: Continuous symmetry and a conjecture, Zeitschrift f¨ur Physik B Con- +densed Matter 35 (1979) 207. +[65] A. J. McKane and M. Stone, Localization as adn Alternative to Goldstone’s Theorem, Annals of Physics 131 +(1981) 36–55. +[66] Shinobu Hikami, Anderson localization in a non-lienar σ-model representation, Physical Review B 24 (1981) +2671–2679. +[67] K. G. Wilson, Quantum Field Theory Models in Less Than 4 Dimensions, Phys. Rev. D 7 (1973) 2911. +[68] S. L. Sondhi, S. M. Girvin, J. P. Carini, D. Shahar, Continuous quantum phase transitions, Rev. Mod. Phys. 69 +(1997) 315. +[69] S. Sachdev, Quantum Phase Transitions, 2nd Edition, Cambridge University Press, Cambridge, UK, 2011. +[70] E. Fradkin, L. Susskind, Order and disorder in gauge systems and magnets, Phys. Rev. D 17 (1978) 2637. +[71] J. B. Kogut, An introduction to lattice gauge theory and spin systems, Rev. Mod. Phys. 51 (1979) 659. +[72] P. M. Chaikin, T. C. Lubensky, Principles of Condensed Matter Physics, Cambridge University Press, Cambridge, +U.K., 1995. +[73] E. Ardonne, P. Fendley, E. Fradkin, Topological order and conformal quantum critical points, Annals of Physics +310 (2004) 493. +[74] A. Perelomov, Generalized Coherent States and Their Applications, Springer-Verlag, Berlin, 1986. +[75] E. Fradkin, M. Stone, Topological terms in one- and two-dimensional quantum Heisenberg antiferromagnets, +Phys. Rev. B 38 (1988) 7215(R). +[76] P. B. Wiegmann, Multivalued functionals and geometrical approach for quantization of relativistic particles and +strings, Nuclear Physics B 323 (1989) 311. +[77] M. V. Berry, Quantal Phase Factors Accompanying Adiabatic Changes, Proceedings of the Royal Society of +London A 392 (1984) 45. +117 + +[78] B. Simon, Holonomy, the Quantum Adiabatic Theorem, and Berry’s Phase, Phys. Rev. Lett. 51 (1983) 2167. +[79] D. Dombre, N. Read, Absence of the Hopf invariant in the long-wavelength action of two-dimensional antiferro- +magnets, Phys. Rev. B 38 (1988) 7181. +[80] F. D. M. Haldane, O(3) non-linear σ model and the topological distinction between integer and half-integer-spin +antiferromagnets in two dimensions, Phys. Rev. Lett. 61 (1988) 1029. +[81] S. Chakravarty, B. I. Halperin, D. R. Nelson, Low-temperature behavior of two-dimensional quantum antiferro- +magnets, Phys. Rev. Lett. 60 (1988) 1057. +[82] N. D. Mermin, The topological theory of defects in ordered media, Rev. Mod. Phys. 51 (1979) 591. +[83] A. A. Abrikosov, On the magnetic properties of superconductors of the second group, Soviet Physics JETP 5 +(1957) 1174, Zh. Ekzp. i Teor. Fiz. 32, 1442 (1957). +[84] J. M. Kosterlitz, D. J. Thouless, Order, metastability and phase transitions in two-dimensional systems, J. Phys. +C: Solid State Phys. 6 (1973) 1181. +[85] J. M. Kosterlitz, The critical properties of the two-dimensional XY model, J. Phys. C: Solid State Phys. 7 (1974) +1046. +[86] V. L. Berezinskii, Destruction of Long-Range Order in One-Dimensional and Two-Dimensional Systems having +a Continuous Symmetry Group I. Classical Systems, Soviet Physics JETP 32 (1971) 493. +[87] V. L. Berezinskii, Destruction of Long-Range Order in One-Dimensional and Two-Dimensional Systems having +a Continuous Symmetry Group II. Quantum Systems, Soviet Physics JETP 34 (1972) 610. +[88] J. V. Jos´e, L. P. Kadanoff, S. Kirkpatrick, D. R. Nelson, Renormalization, vortices, and symmetry-breaking per- +turbations in the two-dimensional planar model, Phys. Rev. B 16 (1977) 1217. +[89] B. I. Halperin, D. R. Nelson, Theory of Two-Dimensional Melting, Physical Review Letters 41 (1978) 121. +[90] A. P. Young, Melting and the vector Coulomb gas in two dimensions, Physical Review B 19 (1979) 1855. +[91] J. Toner, D. R. Nelson, Smectic, cholesteric, and Rayleigh-Benard order in two dimensions, Physical Review B +23 (1981) 316. +[92] D. R. Nelson, J. Toner, Bond-orientational order, dislocation loops, and melting of solids and smectic-A liquid +crystals, Physical Review B 24 (1981) 363. +[93] S. Coleman, Aspects of Symmetry, Cambridge University Press, Cambridge U.K., 1985. +[94] R. Rajaraman, Solitons and Instantons, North-Holland, Amsterdam, The Netherlands, 1985. +[95] H. B. Nielsen, P. Olesen, Vortex-line models for dual strings, Nuclear Physics B 61 (1973) 45. +[96] G. ’t Hooft, Symmetry Breaking through Bell-Jackiw Anomalies, Physical Review Letters 37 (1976) 8. +[97] A. M. Polyakov, Quark confinement and topology of gauge theories, Nucl. Phys. B 120 (1977) 429. +[98] A. M. Polyakov, Gauge Fields and Strings, Harwood Academic Publishers, London, 1987. +[99] C. G. Callan, R. F. Dashen, D. J. Gross, The structure of the gauge theory vacuum, Physics Letters B 63 (1976) +334. +[100] A. O. Caldeira, A. J. Leggett, Quantum Tunneling in a Dissipative System, Annals of Physics 149 (1983) 374–456. +[101] A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. Fisher, A. Garg, W. Zwerger, Dynamics of the dissipatuve two-state +system, Reviews of Modern Physics 59 (1987) 1–85. +[102] P. Di Francesco, P. Mathieu, D. S´en´echal, Conformal Field Theory, Springer-Verlag, Berlin, 1997. +[103] P. Ginsparg, Applied Conformal Field Theory, in: E. Br´ezin, J. Zinn-Justin (Eds.), Champs, Cordes et Ph´enom´enes +Critiques/ Fields, Strings and Critical Phenomena, Proceedings of the Les Houches Summer School 1988, Session +XLIX, Elsevier Science Publishers B.V., Amsterdam, The Netherlands, 1989, pp. 1–168. +[104] J. Polchinski, String Theory, Cambridge University Press, Cambridge U.K., 1998. +[105] L. P. Kadanoff, Lattice Coulomb gas representations of two-dimensional problems, Journal of Physics A: Mathe- +matical and Theoretical 11 (1978) 1399. +[106] B. Nienhuis, Two-dimensional critical phenomena and the Coulomb gas, in: C. Domb, M. Green, J. L. Lebowitz +(Eds.), Phase Transitions and Critical Phenomena, Vol. 11, Academic Press, London, U. K., 1987, p. 1. +[107] N. D. Mermin, H. Wagner, Absence of Ferromagnetism or Antiferromagnetism in One- or Two-Dimensional +Isotropic Heisenberg Models, Phys. Rev. Lett. 17 (1966) 1133–1136. +[108] P. C. Hohenberg, Existence of Long-Range Order in One and Two Dimensions, Phys. Rev. 158 (1967) 383–386. +[109] S. Coleman, There are no Goldstone bosons in two dimensions, Communications in Mathematical Physics 31 +(1973) 259–264. +[110] F. A. Schaposnik, Pseudoparticles and confinement in the two-dimensional Abelian Higgs model, Physical Review +D 18 (1978) 1183. +[111] G. ’t Hooft, Computation of the quantum effects due to a four-dimensional pseudoparticle, Phys. Rev. D 14 (1976) +3432–3450. +[112] H. Georgi, S. Glashow, Unity of all elementary-particle forces, Physical Review Letters 32 (1974) 438. +[113] A. Polyakov, Compact gauge fields and the infrared catastrophe, Physics Letters B 59 (1975) 82 – 84. +[114] P. Dirac, Quantised singularities in the electromagnetic field, Proc. Roy. Soc. London 133 (1931) 60. +[115] F. D. M. Haldane, ‘θ’ physics’ and quantum spin chains, J. Appl. Phys. 57 (1985) 3359. +[116] H. Bethe, Theory of metals. I. Eigenvalues and eigenfunctions of the linear atomic chain, Z. Physik 71 (1931) +205. +[117] C. N. Yang, C. P. Yang, Thermodynamics of a One-Dimensional System of Bosons with Repulsive Delta-Function +Interaction, J. Math. Phys. 10 (1969) 1115. +[118] I. Affleck, F. D. M. Haldane, Critical theory of quantum spin chains, Phys. Rev. B 36 (1987) 5291. +[119] E. Witten, Non-Abelian Bosonization in Two Dimensions, Comm. Math. Phys. 92 (1984) 455. +[120] V. G. Knizhnik, A. B. Zamolodchikov, Current Algebra and Wess-Zumino Model in Two Dimensions, Nuclear +Physics B 247 (1984) 83. +[121] I. Affleck, T. Kennedy, E. H. Lieb, H. Tasaki, Valence Bond Ground States in Isotropic Quantum Antiferromag- +nets, Comm. Math. Phys. 115 (1988) 477. +[122] M. Hagiwara, K. Katsumata, I. Affleck, B. I. Halperin, J. P. Renard, Observation of S = 1/2 degrees of Freedom +in an S = 1 Linear-Chain Heisenberg Antiferromagnet, Physical Review Letters 65 (1990) 3181. +[123] H. A. Kramers, G. H. Wannier, Statistics of the Two-Dimensional Ferromagnet. Part I, Phys. Rev. 60 (1941) 252. +[124] F. J. Wegner, Duality in Generalized Ising Models and Phase Transitions without Local Order Parameters, J. Math. +Phys. 12 (1971) 2259. +[125] S. Elitzur, R. B. Pearson, J. Shigemitsu, Phase structure of discrete Abelain spin and gauge systems, Physical +118 + +Review D 19 (1979) 3698. +[126] A. Ukawa, P. Windey, A. H. Guth, Dual variables for lattice gauge theories and the phase structure of Z(N) +systems, Physical Review D 21 (1980) 1013. +[127] F. Haldane, Luttinger liquid thepry of one-dimensional quantum fluids. I. Properties of the Luttinger model and +their extension to the general 1D interacting spinless Fermi gas, J. Phys. C: Solid State Phys. 14 (1981) 2585. +[128] D. C. Mattis, E. H. Lieb, Exact Solution of a Many-Fermion System and its Associated Boson Field, Journal of +Mathematical Physics 6 (1965) 304. +[129] A. Luther, V. J. Emery, Backward Scattering in the One-Dimensional Electron Gas, Physical Review Letters 33 +(1974) 589. +[130] S. Mandelstam, Soliton operators for the quantized sine-Gordon equation, Physical Review D 11 (1975) 3026. +[131] S. Coleman, Quantum sine-Gordon equation as the massive Thirring model, Phys. Rev. D 11 (1975) 2088. +[132] D. J. Gross, A. Neveu, Dynamical symmetry breaking in asymptotically free field theories, Phys. Rev. D 10 (1974) +3235. +[133] S. E. Adler, The Axial-Vector Vertex in Spinor Electrodynamics, Physical Review 177 (1969) 2426. +[134] J. S. Bell, R. Jackiw, The PCAC Puzzle: π0 → γγ in the σ-model, Nuovo Cimento A 60 (1969) 47. +[135] X. L. Qi, T. L. Hughes, S. C. Zhang, Topological field theory of time-reversal invariant insulators, Phys. Rev. B +78 (2008) 195424. +[136] J. Schwinger, Field Theory Commutators, Phys. Rev. Lett. 3 (1959) 296. +[137] E. Witten, Some Properties of the ( ¯ψψ)2 Model in Two Dimensions, Nucl. Phys. B 142 (1978) 285. +[138] W. P. Su, J. R. Schrieffer, A. J. Heeger, Solitons in Polyacetylene, Phys. Rev. Lett. 42 (1979) 1698. +[139] R. Jackiw, J. R. Schrieffer, Solitons with fermion number 1/2 in condensed matter and relativistic field theories, +Nuclear Physics B 190 (1981) 253. +[140] R. Jackiw, C. Rebbi, Solitons with fermion number 1/2, Physical Review D 13 (1976) 3398. +[141] J. Goldstone, F. Wilczek, Fractional Quantum Numbers on Solitons, Phys. Rev. Lett. 47 (1981) 986. +[142] H. Takayama, Y. R. Lin-Liu, K. Maki, Continuum model for solitons in polyacetylene, Phys. Rev. B 21 (1980) +2388–2393. +[143] D. K. Campbell, A. R. Bishop, Solitons in polyacetylene and relativistic-field-theory models, Phys. Rev. B 24 +(1981) 4859–4862. +[144] D. Campbell, A. Bishop, Soliton excitations in polyacetylene and relativistic field theory models, Nuclear Physics +B 200 (1982) 297–328. +[145] J. E. Hirsch, E. Fradkin, Effect of quantum fluctuations on the peierls instability: A monte carlo study, Phys. Rev. +Lett. 49 (1982) 402–405. +[146] E. Fradkin, J. E. Hirsch, Phase diagram of one-dimensional electron-phonon systems. I. The Su-Schrieffer-Heeger +model, Phys. Rev. B 27 (1983) 1680. +[147] M. F. Atiyah, V. K. Patodi, I. M. Singer, Spectral Asymmetry and Riemaniann Geometry. I, Math. Proc. Cambridge +Philos. Soc. 77 (1975) 43. +[148] P. Dirac, The Principles of Quantum Mechanics, Oxford University Press, Oxford, U.K., 1930. +[149] L. D. Landau, E. M. Lifshitz, Quantum Mechanics: Non-Relativistic Theory, 1st Edition, Vol. Volume 3 of Course +of Theoretical Physics, Pergamon Press, Oxford, U.K., 1957. +[150] S. Weinberg, The Quantum Theory Of Fields, 1st Edition, Vol. I-III, Cambridge University Press, Cambridge, +U.K., 2005. +[151] T. T. Wu, C. N. Yang, Dirac monopole without strings: Monopole harmonics, Nuclear Physics B 107 (1976) 365. +[152] E. Fradkin, L. P. Kadanoff, Disorder variables and para-fermions in two-dimensional statistical mechanics, Nu- +clear Physics B 170 (1980) 1. +[153] J. M. Leinaas, J. Myrheim, On the theory of identical particles, Il Nuovo Cimento B (1971-1996) 37 (1977) 1–23. +[154] M. G. G. Laidlaw, C. M. DeWitt, Feynman Functional Integrals for Systems of Indistinguishable Particles, Phys- +ical Review D 3 (1971) 1375–1378. +[155] Y. Aharonov, D. Bohm, Significance of Electromagnetic Potentials in the Quantum Theory, Phys. Rev. 115 (1959) +485–491. +[156] F. Wilczek, Magnetic Flux, Angular Momentum, and Statistics, Phys. Rev. Lett. 48 (1982) 1144. +[157] F. Wilczek, Quantum Mechanics of Fractional-Spin Particles, Phys. Rev. Lett. 49 (1982) 957–959. +[158] A. S. Goldhaber, Connection of Spin and Statistics for Charge-Monopole Composites, Phys. Rev. Lett. 36 (1976) +1122–1125. +[159] D. Finkelstein, J. Rubinstein, Connection between Spin, Statistics, and Kinks, Journal of Mathematical Physics 9 +(1968) 1762–1779. +[160] T. Skyrme, A Unified Theory of Mesons and Baryons, Nuclear Physics 31 (1962) 556–569. +[161] Y.-S. Wu, General Theory for Quantum Statistics in Two Dimensions, Phys. Rev. Lett. 52 (1984) 2103–2106. +[162] F. Wilczek, A. Zee, Linking numbers, spin, and statistics of solitons, Phys. Rev. Lett. 51 (1983) 2250. +[163] S. Deser, R. Jackiw, S. Templeton, Three-Dimensional Massive Gauge Theories, Physical Review Letters 48 +(1982) 975. +[164] S. Deser, R. Jackiw, S. templeton, Topologically Massive Gauge Theories, Annals of Physics 140 (1982) 372–411. +[165] E. Witten, Quantum Field Theory and the Jones Polynomial, Commun. Math. Phys. 121 (1989) 351. +[166] G. T. Horowitz, Exactly Soluble Diffeomorphism Invaraint Theories, Communications in Mathematical Physics +125 (1989) 417. +[167] J. Grundberg, T. H. Hansson, A. Karlhede, On Polyakov’s Spin Factors, Nuclear Physics B 347 (1990) 420–440. +[168] A. M. Polyakov, Fermi-Bose Transmutations Induced by Gauge Fields, Mod. Phys. Lett. A 3 (1988) 325. +[169] A. +Y. +Kitaev, +Fault-tolerant +quantum +computation +by +anyons, +Annals +of +Physics +303 +(2003) +2, +(arXiv:quant-ph/9707021). +[170] S. Das Sarma, M. Freedman, C. Nayak, S. H. Simon, A. Stern, Non-abelian anyons and topological quantum +computation, Rev. Mod. Phys. 80 (2008) 1083. +[171] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, A. K. Geim, The electronic properties of graphene, +Rev. Mod. Phys. 81 (2009) 109–162. +[172] X. L. Qi, S. C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83 (2011) 1057. +[173] L. Susskind, Lattice Fermions, Physical Review D 16 (1977) 3031. +[174] H. B. Nielsen, M. Ninomiya, Absence of neutrinos on a lattice (I). Proof by homotopy theory, Nuclear Physics B +119 + +185 (1981) 20–40. +[175] H. B. Nielsen, M. Ninomiya, Absence of neutrinos on a lattice (II). Intiuitive topological proof, Nuclear Physics +B 193 (1981) 173–194. +[176] H. B. Nielsen, M. Ninomiya, The Adler-Bell-Jackiw Anomaly and Weyl fermions in a Crystal, Physics Letters B +130 (1983) 389. +[177] D. Friedan, A Proof of the Nielsen-Ninomiya Theorem, Communications in Mathematical Physics 85 (1982) +481–490. +[178] G. W. Semenoff, Condensed-Matter Simulation of a Three-Dimensional Anomaly, Physical Review Letters 53 +(1984) 2449. +[179] D. J. Thouless, M. Kohmoto, M. P. Nightingale, M. den Nijs, Quantized Hall Conductance in a Two-Dimensional +Periodic Potential, Physical Review Letters 49 (1982) 405. +[180] I. Affleck, J. B. Marston, Large-N limit of the Heisenberg-Hubbard Model: Implications for High-Tc Supercon- +ductors, Phys. Rev. B 37 (1988) 3774. +[181] X. G. Wen, F. Wilczek, A. Zee, Chiral spin states and superconductivity, Phys. Rev. B 39 (1989) 11413–11423. +[182] F. Bloch, ¨Uber der quantenmechanik der elektronen in kristallgittern, Zeitschrift f¨ur physik 52 (1929) 555–600. +[183] C. Kittel, Introduction to Solid State Physics, 1st Edition, John Wiley & Sons, New York, New York, 1996. +[184] N. Ashcroft, D. N. Mermin, Solid State Physics, Holt, Reinhart and Winston, New York, New York, 1976. +[185] B. Bradlyn, L. Elcoro, J. Cano, M. G. Vergniory, Z. Wang, C. Felser, M. I. Aroyo, B. A. Bernevig, Topological +quantum chemistry, Nature 547 (2017) 298–305. +[186] J. Cano, B. Bradlyn, Band Representations and Topological Quantum Chemistry, Annual Review of Condensed +Matter Physics 12 (2021) 225–246. +[187] M. Kohmoto, Topological invariant and the quantization of the Hall conductance, Annals of Physics 160 (1985) +343. +[188] A. P. Schnyder, S. Ryu, A. Furusaki, A. W. W.Ludwig, Classification of topological insulators and superconductors +in three spatial dimensions, Phys. Rev. B 78 (2008) 195125. +[189] A. Kitaev, Periodic table for topological insulators and superconductors, in: M. Feigelman (Ed.), Advances in +Theoretical Physics: Landau Memorial Conference, Vol. 1134, American Institute of Physics, AIP Conference +Proceedings, College Park, Maryland, 2009, p. 22. +[190] J. E. Moore, L. Balents, Topological invariants of time-reversal-invariant band structures, Phys. Rev. B 75 (2007) +121306. +[191] E. Fradkin, M. Kohmoto, Quantized Hall effect and geometric localization of electrons on lattices, Phys. Rev. B +35 (1987) 6017–6023. +[192] D. R. Hofstadter, Energy levels and wave functions of Bloch electrons in rational and irrational magnetic fields, +Physical Review B 14 (1876) 2239. +[193] P. C. Martin, Measurements and Correlation Functions, Gordon & Breach, Philadelphia, Pennsylvania, 1967. +[194] P. Nozi`eres, D. Pines, The Theory of Quantum Liquids: Normal Fermi Liquids, W. A. Benjamin Inc., New York, +New York, 1966. +[195] S. Ryu, C. Mudry, C.-Y. Hou, C. Chamon, Masses in graphene-like two-dimensional electronic systems: Topo- +logical defects in order parameters and their fractional exchange statistics, Physical Review B 80 (2009) 205319. +[196] F. D. M. Haldane, Model for a Quantum Hall Effect without Landau Levels: Condensed-Matter Realization of the +“Parity Anomaly”, Phys. Rev. Lett. 61 (1988) 2015–2018. +[197] X. Chen, Z.-C. Gu, Z.-X. Liu, X.-G. Wen, Symmetry-Protected Topological Orders in Interacting Bosonic Sys- +tems, Science 338 (2012) 1604–1606. +[198] T. Senthil, Symmetry-protected topological phases of quantum matter, Annual Review of Condensed Matter +Physics 6 (2015) 299–324. +[199] E. Witten, Fermion path integrals and topological phases, Reviews of Modern Physics 88 (2016) 035001. +[200] A. N. Redlich, Gauge Noninvariance and Parity Nonconservation of Three-Dimensional Fermions, Physical Re- +view Letters 52 (1984) 18. +[201] A. N. Redlich, Parity violation and gauge noninvariance of the effective gauge field action in three dimensions, +Phys. Rev. D 29 (1984) 2366. +[202] K. Fujikawa, Path-Integral Measure for gauge-Invariant Fermion Theories, Physical Review Letters 42 (1979) +1195. +[203] R. E. Gamboa Sarav´ı, M. A. Muschietti, F. A. Schaposnik, J. E. Solomin, Chiral Symmetry and Functional +Integral, Annals of Physics 157 (1984) 360. +[204] G. E. Volovik, Zeros in the fermion spectrum in superfluid systems as diabolical points, Journal of Experimental +and Theoretical Physics Letters 46 (1987) 98. +[205] V. M. Yakovenko, Chern-Simons Terms and n Field in Haldane’s Model for the Quantum Hall Effect without +Landau Levels, Physical Review Letters 65 (1990) 251. +[206] M. F. Golterman, K. Jansen, D. B. Kaplan, Chern-Simons currents and chiral fermions on the lattice, Physics +Letters B 301 (1993) 219. +[207] X.-L. Qi, Y.-S. Wu, S.-C. Zhang, Topological quantization of the spin Hall effect in two-dimensional paramagnetic +semiconductors, Physical Review B 74 (2006) 085308. +[208] N. Seiberg, T. Senthil, C. Wang, E. Witten, A Duality Web in 2+1 Dimensions and Condensed Matter Physics, +Annals of Physics 374 (2016) 395 – 433. +[209] L. Fu, C. L. Kane, Topological insulators with inversion symmetry, Phys. Rev. B 76 (2007) 045302. +[210] C. L. Kane, E. J. Mele, Z2 Topological Order and the Quantum Spin Hall Effect, Phys. Rev. Lett. 95 (2005) +146802. +[211] B. A. Bernevig, T. L. Hughes, S. C. Zhang, Quantum Spin Hall Effect and Topological Phase Transition in HgTe +Quantum Wells, Science 314 (2006) 1757. +[212] J. Gooth, B. Bradlyn, S. Honnali, C. Schindler, N. Kumar, J. Noky, Y. Qi, C. Shekhar, Y. Sun, Z. Wang, B. A. +Bernevig, C. Felser, Axionic charge-density wave in the Weyl semimetal (TaSe4)2I, Nature 575 (2019) 315–319. +[213] A. M. Essin, J. E. Moore, D. Vanderbilt, Magnetic Polarizability and Axion Electrodynamics in Crystalline Insu- +lators, Physical Review Letters 102 (2009) 146805. +[214] P. Hosur, S. Ryu, A. Vishwanath, Chiral topological insulators, superconductors, and other competing orders in +three diemsnions, Physical Review B 81 (2010) 045120. +120 + +[215] F. Wilczek, Two applications of axion electrodynamics, Phys. Rev. Lett. 58 (1987) 1799–1802. +[216] Michael E. Peskin and Daniel V. Schroeder, An Introduction to Quantum Field Theory, The Advanced Book +Program, Perseus Books Publishing L.L.C., Reading, Massachusetts, 1995. +[217] E. Witten, Dyons of charge eθ/2π, Physics Letters B 86 (1979) 283. +[218] C. G. Callan, J. A. Harvey, Anomalies and Fermion Zero Modes on Strings and Domain Walls, Nucl. Phys. B 250 +(1985) 427. +[219] D. Hsieh, Y. Xia, D. Qian, L. Wray, J. H. Dil, F. Meier, J. Osterwalder, L. Patthey, J. G. Checkelsky, N. P. Ong, +A. V. Fedorov, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. Cava, M. Z. Hasan, A tunable topological insulator +in the spin helical Dirac transport regime, Nature 460 (2009) 1101–1105. +[220] K. von Klitzing, G. Dorda, M. Pepper, New Method for High-Accuracy Determination of the Fine-Structure +Constant Based on Quantized Hall Resistance, Physical Review Letters 45 (1980) 494. +[221] D. C. Tsui, H. L. Stormer, A. C. Gossard, Two-Dimensional Magnetotransport in the Extreme Quantum Limit, +Physical Review Letters 48 (1982) 1559. +[222] X. Du, I. Skachko, F. Duerr, A. Luican, E. Y. Andrei, Fractional quantum Hall effect and insulating phase of Dirac +electrons in graphene, Nature 462 (2009) 192–195. +[223] A. A. Zibrov, C. Kometter, H. Zhou, E. M. Spanton, T. Taniguchi, K. Watanabe, M. P. Zaletel, A. F. Young, +Tunable interacting composite fermion phases in a half-filled bilayer-graphene Landau level, Nature 549 (2017) +360–364. +[224] L. Landau, Diamagnetismus der Metalle, Zeitschrift f¨ur Physik 64 (1930) 629–637. +[225] Q. Niu, D. J. Thouless, Y.-S. Wu, Quantized Hall conductance as a topological invariant, Physical Review B 31 +(1985) 3372. +[226] R. B. Laughlin, Quantized Hall conductivity in two dimensions, Physical Review B 23 (1981) 5632–5633. +[227] H. Levine, S. B. Libby, A. Pruisken, Electron Delocalization by a Magnetic Field in Two Dimensions, Physical +Review Letters 51 (1983) 1915. +[228] A. Pruisken, On localization in the theory of the quantized hall effect: A two-dimensional realization of the +θ-vacuum, Nuclear Physics B 235 (1984) 277. +[229] S. Kievelson, M. Ro˘cek, Consequences of Gauge Invariance for Fractionally Charged Quasi-Particles, Physics +Letters B 85 (156). +[230] J. K. Jain, Incompressible quantum Hall states, Physical Review B 40 (1989) 8079. +[231] J. K. Jain, Composite-fermion approach for the fractional quantum Hall effect, Physical Review Letters 63 (1989) +199. +[232] B. I. Halperin, P. A. Lee, N. Read, Theory of the half-filled Landau level, Phys. Rev. B 47 (1993) 7312. +[233] J. Nakamura, S. Liang, G. C. Gardner, M. J. Manfra, Direct observation of anyonic braiding statistics, Nature +Physics 16 (2020) 931–936. +[234] B. I. Halperin, Statistics of Quasiparticles and the Hierarchy of Fractional Quantized Hall States, Physical Review +Letters 52 (1984) 1583. +[235] D. Arovas, J. R. Schrieffer, F. Wilczek, Fractional Statistics and the Quantum Hall Effect, Phys. Rev. Lett. 53 +(1984) 722. +[236] J. Fr¨ohlich, A. Zee, Large-scale physics of the quantum Hall Fluid, Nuclear Physics B 364 (1991) 517. +[237] V. Pasquier, F. Haldane, A dipole interpretation of the ν = 1/2 state, Nuclear Physics B 516 (1998) 719. +[238] N. Read, Lowest-Landau-level theory of the quantum Hall effect: The Fermi-liquid-like state of bosons at filling +factor one, Phys. Rev. B 58 (1998) 16262. +[239] L. Susskind, The Quantum Hall Fluid and Non-Commutative Chern-Simons Theory, unpublished (2001). +arXiv:arXiv:hep-th/0101029. +[240] A. P. Polychronakos, Quantum Hall States as matrix Chern-Simons theory, Journal of High Eenergy Physics 2001 +(2001) 011. +[241] E. Fradkin, V. Jejjala, R. G. Leigh, Non-commutative Chern-Simons for the Quantum Hall system and Duality, +Nuclear Physics B 642 (2002) 483–500. +[242] Z. Dong, T. Senthil, Noncommutative field theory and composite Fermi liquids in some quantum Hall systems, +Phys. Rev. B 102 (2020) 205126. +[243] H. Goldman, T. Senthil, Lowest Landau level theory of the bosonic Jain states, Phys. Rev. B 105 (2022) 075130. +[244] S. C. Zhang, T. H. Hansson, S. Kivelson, Effective-Field-Theory Model for the Fractional Quantum Hall Effect, +Phys. Rev. Lett. 62 (1989) 82. +[245] F. Haldane, E. H. Rezayi, Periodic Laughlin-Jastrow wave functions for the fractional quantized Hall effect, +Physical Review B 31 (1985) 2529. +[246] X. G. Wen, Q. Niu, Ground-state degeneracy of the fractional quantum Hall states in the presence of a random +potential and on high-genus Riemann surfaces, Phys. Rev. B 41 (1990) 9377. +[247] A. L´opez, E. Fradkin, Fractional quantum Hall effect and Chern-Simons gauge theories, Phys. Rev. B 44 (1991) +5246. +[248] A. L´opez, E. Fradkin, Response Functions and Spectrum of Collective Excitations of Fractional Quantum Hall +Effect Systems, Physical Review B 47 (1993) 7080. +[249] W. Kohn, Cyclotron Resonance and de Haas-van Alphen Oscillations of an Interacting Electron Gas, Physical +Review 123 (1961) 1242. +[250] A. L´opez, E. Fradkin, Universal structure of the edge states of the fractional quantum Hall states, Physical Review +B 59 (1999) 15323. +[251] X.-G. Wen, A. Zee, Classification of Abelian quantum Hall states and matrix formulation of topological fluids, +Physical Review B 46 (1992) 2290. +[252] X. G. Wen, Topological Orders and Edge Excitations in Fractional Quantum Hall States, Advances in Physics 44 +(1995) 405. +[253] R. de Picciotto, M. Reznikov, M. Heiblum, V. Umansky, G. Bunin, D. Mahalu, Direct observation of a fractional +charge, Nature 389 (1997) 162. +[254] L. Saminadayar, D. C. Glattli, Y. Jin, B. Etienne, Observation of the e/3 Fractionally Charged Laughlin Quasipar- +ticle, Physical Review Letters 79 (1997) 2526. +[255] J. A. Hertz, Quantum critical phenomena, Phys. Rev. B 14 (1976) 1165–1184. +[256] A. J. Millis, Effect of a nonzero temperature on quantum critical points in itinerant fermion systems, Physical +121 + +Review B 48 (1993) 7183–7196. +[257] C. M. Varma, P. B. Littlewood, S. Schmitt-Rink, E. Abrahams, A. E. Ruckenstein, Phenomenology of the normal +state of Cu-O high-temperature superconductors, Phys. Rev. Lett. 63 (1989) 1996–1999. +[258] J. Polchinski, Low-energy dynamics of the spinon-gauge system, Nuclear Physics B 422 (1994) 617–633. +[259] R. Shankar, Renormalization-group approach to interacting fermions, Reviews of Modern Physics 66 (1994) 129– +192. +[260] J. Polchinski, Effective Field Theory and the Fermi Surface, in: J. A. Harvey, J. Polchinski (Eds.), Theoretical +Advanced Study Institute (TASI 92): From Black Holes and Strings to Perticles, World-Scientific Publishing Co., +Singapore, 1993, pp. 235–276, proceedings of TASI 1992, Boulder (Colorado, USA), June 1-26 (1992). +[261] C. Nayak, F. Wilczek, Renormalization group approach to low temperature properties of a non-Fermi liquid metal, +Nuclear Physics B 430 (1994) 534–562. +[262] C. Nayak, F. Wilczek, Non-Fermi liquid fixed point in 2 + 1 dimensions, Nuclear Physics B 417 (1994) 359–373. +[263] J. M. Maldacena, Large N limit of superconformal field theories and supergravity, Advances in Theoretical and +Mathematical Physics 2 (1998) 231. +[264] T. Faulkner, N. Iqbal, H. Liu, J. McGreevy, D. Vegh, Strange Metal Transport Realized by Gauge/Gravity Duality, +Science 329 (2010) 1043–1047. +[265] S. Fubini, Vertex Operators and the Quantum Hall Effect, Modern Physics Letters A 6 (1991) 347–358. +[266] G. Moore, N. Read, Non-Abelions in the Fractional Quantum Hall Effect, Nuclear Physics B 360 (1991) 362. +[267] G. Moore, N. Seiberg, Taming the Conformal Zoo, Phys. Lett. B 220 (1989) 422. +[268] G. Moore, N. Seiberg, Classical and Quantum Conformal Field Theory, Commun. Math. Phys. 123 (1989) 177. +[269] M. H. Freedman, A. Kitaev, Z. Wang, Simulation of topological field theories by quantum computers, Commun. +Math. Phys. 227 (2002) 587. +[270] M. H. Freedman, A. Kitaev, M. J. Larsen, Z. Wang, Topological Quantum Computation, Comm. Math. Phys. 227 +(2002) 605. +[271] J. Preskill, Topological Quantum Computation, Lecture Notes for Physics 219: Quantum Computation, chapter +9; unpublished; Caltech (2004). +URL http://www.theory.caltech.edu/~preskill/ph219/topological.pdf +[272] R. Willett, J. P. Eisenstein, H. L. St¨ormer, D. C. Tsui, A. C. Gossard, J. H. English, Observation of an even- +denominator quantum number in the fractional quantum Hall effect, Physical Review Letters 59 (1987) 1776. +[273] J. P. Eisenstein, R. Willett, H. L. St¨ormer, D. C. Tsui, A. C. Gossard, J. H. English, Collapse of the Even- +Denominator Fractional Quantum Hall Effcet in Tilted Fields, Physical Review Letters 61 (1988) 997. +[274] W. Pan, J. S. Xia, H. L. St¨ormer, D. C. Tsui, C. Vicente, E. D. Adams, N. S. Sullivan, L. N. Pfeiffer, K. W. +Baldwin, K. W. West, Experimental studies of the fractional quantum Hall effect in the first excited Landau level, +Phys. Rev. B 77 (2008) 075307. +[275] M. Greiter, X.-G. Wen, F. Wilczek, Paired hall State at Half Filling, Physical Review Letters 66 (1991) 3205. +[276] M. Greiter, X. G. Wen, F. Wilczek, Paired Hall States, Nuclear Physics B 374 (1992) 567–614. +[277] X. G. Wen, Non-Abelian statistics in the Fractional quantum Hall States, Physical Review Letters 66 (1991) 802. +[278] W. Kohn, J. M. Luttinger, New Mechanism for Superconductivity, Physical Review Letters 15 (1965) 524. +[279] A. V. Chubukov, Kohn-Luttinger effect and the instability of a two-dimensional repulsive Fermi liquid at T = 0, +Physical Review B 48 (1993) 1097. +[280] S. Raghu, S. A. Kivelson, Superconductivity from repulsive interactions in the two-dimensional electron gas, +Physical Review B 83 (2011) 094518. +[281] K. Park, V. Melik-Alaverdian, N. E. Bonesteel, J. K. Jain, Possibility of p-wave pairing of composite fermions at +ν = 1/2, Physical Review B 58 (1998) R10167. +[282] B. I. Halperin, Theory of the quantized Hall conductance, hevetica Physica Acta 56 (1983) 75–102. +[283] R. H. Morf, Transition from quantum Hall to compressible states in the second Landau level: New light on the +ν = 5/2 enigma, Physical Review Letters 80 (1998) 1505. +[284] C. Nayak, F. Wilczek, 2n-quasihole states realize 2n−1-dimensional spinor braiding statistics in paired quantum +Hall states, Nuclear Physics B 479 (1996) 529–553. +[285] N. Read, D. Green, Paired states of fermions in two dimensions with breaking of parity and time-reversal symme- +tries and the fractional quantum Hall effect, Physical Review B 61 (2000) 10267. +[286] G. E. Volovik, An analog of the quantum Hall effect in a superfluid 3He film, Soviet Physics JETP 67 (1988) +1804. +[287] R. Jackiw, P. Rossi, Zero modes of the vortex-fermion system, Nuclear Physics B 190 (1981) 681. +[288] E. J. Weinberg, Index calculations for the fermion-vortex system, Phys. Rev. D 24 (1981) 2669. +[289] Y. Nishida, L. Santos, C. Chamon, Topological superconductors as nonrelativistic limits of Jackiw-Rossi and +Jackiw-Rebbi models, Phys. Rev. B 82 (2010) 144513. +[290] D. A. Ivanov, Non-abelian statistics of half-quantum vortices in p-wave superconductors, Phys. Rev. Lett. 86 +(2001) 268. +[291] M. Stone, S.-B. Chung, Fusion rules and vortices in px + ipy superconductors, Phys. Rev. B 73 (2006) 014505. +[292] N. Read, E. Rezayi, Beyond paired quantum Hall states: Parafermions and incompressible states in the first excited +Landau level, Physical Review B 59 (1999) 8084. +[293] A. B. Z. V. A. Fateev, Nonlocal(parafermion) currents in two-dimensional conformal quantum field theory and +self-dual critical points in ZN-symmetric statistical systems, Soviet Physics JETP 62 (1985) 215. +[294] F. C. Alcaraz, R. K¨oberle, Hidden parafermions in Z(N) theories, Physical Review D 24 (1981) 1652. +[295] W. Pan, H. L. St¨ormer, D. C. Tsui, L. N. Pfeiffer, K. W. Baldwin, K. W. West, Fractional Quantum Hall Effect of +Composite Fermions, Physical Review B 90 (2003) 016801. +[296] J. S. Xia, W. Pan, C. L. Vicente, E. D. Adams, N. S. Sullivan, H. L. Stormer, D. C. Tsui, L. N. Pfeiffer, K. W. +Baldwin, K. W. West, Electron Correlation in the Second Landau Level: A Competition Between Many Nearly +Degenerate Quantum Phases, Physical Review Letters 93 (2004) 176809. +[297] E. H. Rezayi, N. Read, Non-Abelian quantized Hall states of electrons at filling factors 12/5 and 13/5 in the first +excited Landau level, Physical Review B 79 (2009) 075306. +[298] E. Fradkin, C. Nayak, A. Tsvelik, F. Wilczek, A Chern-Simons effective field theory for the Pfaffian quantum Hall +state, Nuclear Physics B 516 (1998) 704. +[299] E. Fradkin, C. Nayak, K. Schoutens, Landau-Ginzburg theories for non-abelian quantum Hall states, Nuclear +122 + +Physics B 546 (1999) 711. +[300] M. Barkeshli, X.-G. Wen, Effective field theory and projective construction for Zk parafermion fractional quantum +Hall states, Physical Review B 81 (2010) 155302. +[301] H. Goldman, R. Sohal, E. Fradkin, Landau-Gizburg theories of non-Abelian quantum Hall states from non- +Abelian bosonization, Physical Review B 100 (2019) 115111. +[302] H. Goldman, R. Sohal, E. Fradkin, Non-Abelian fermionization and the landscape of quantum Hall phases, Phys. +Rev. B 102 (2020) 195151. +[303] H. Goldman, R. Sohal, E. Fradkin, Composite particle construction of the Fibonacci fractional quantum Hall state, +Phys. Rev. B 103 (2021) 235118. +[304] I. Affleck, Field theory methods and quantum critical phenomena, in: Strings, Fields and Critical Phenomena, +Proceedings of the Les Houches Summer School 1988, Session XLIX, E. Br´ezin and J. Zinn-Justin editors, North- +Holland, Amsterdam, the Netherlands, 1990, p. 563. +[305] A. A. Belavin, A. M. Polyakov, A. B. Zamolodchikov, Infinite Conformal Symmetry in two-Dimensional Quan- +tum Field Theory, Nuclear Physics B 241 (1984) 333. +[306] I. Affleck, Universal Term in the Free Energy at a Critical Point and the Conformal Anomaly, Phys. Rev. Lett. 56 +(1986) 746. +[307] H. Bl¨ote, J. L. Cardy, M. P. Nightingale, Conformal Invariance, the Central Charge, and Universal Finite-Size +Amplitudes at Criticality, Phys. Rev. Lett. 56 (1986) 742. +[308] G. Granger, J. P. Eisenstein, J. L. Reno, Observation of Chiral Heat Transport in the Quantum Hall Regime, +Physical Review Letters 102 (2009) 086803. +[309] A. Bid, N. Ofek, H. Inoue, M. Heiblum, C. L. Kane, V. Umansky, D. Mahalu, Observation of neutral modes in +the fractional quantum Hall regime, Nature 466 (2010) 585–590. +[310] C. L. Kane, M. Fisher, Transmission through barriers and resonant tunneling in an interacting one-dimensional +electron gas, Phys. Rev. B 46 (1992) 15233. +[311] C. Chamon, E. Fradkin, Distinct universal conductances in tunneling to quantum Hall states: The role of contacts, +Phys. Rev. B 56 (1997) 2012. +[312] A. M. Chang, L. N. Pfeiffer, K. W. West, Observation of Chiral Luttinger Behavior in Electron Tunneling into +Fractional Quantum Hall Edges, Physical Review Letters 77 (1996) 2538. +[313] A. M. Chang, Chiral Luttinger liquids at the fractional quantum Hall edge, Reviews of Modern Physics 75 (2004) +1449. +[314] L. A. Cohen, N. L. Samuelson, T. Wang, T. Taniguchi, K. Watanabe, M. P. Zaletel, A. F. Young, Univer- +sal chiral Luttinger liquid behavior in a graphene fractional quantum Hall point contact, unpublished (2022). +arXiv:\protect\vrulewidth0pt\protect\href{http://arxiv.org/abs/2212.01374}{arXiv:2212.01374}. +[315] M. Hashisaka, T. Jonckheere, T. Akiho, S. Sasaki, J. Rech, T. martin, K. Muraki, Andreev reflection of fractional +quantum Hall quasiparticles, Nature Communications 12 (2021) 2794. +[316] N. P. Sandler, C. Chamon, E. Fradkin, Noise measurements and fractional charge in fractional quantum Hall +liquids, Phys. Rev. B 59 (1999) 12521. +[317] F. P. Milliken, C. P. Umbach, R. A. Webb, Indicatiosn of a Luttinger liquid in the fractional quantum Hall regime, +Solid State Communications 97 (1996) 309–313. +[318] S. Roddaro, V. Pellegrini, F. Beltram, G.iorgioBiasiol, L. Sorba, Interedge Strong-to-Weak Scattering Evolution +at a Constriction in the Fractional Quantum Hall Regime, Physical Review Letters 93 (2004) 046801. +[319] S. Roddaro, V. Pellegrini, F. Beltram, Quasi-particle tunneling at a constriction in a fractional quantum Hall state, +Solid State Communications 131 (2004) 565. +[320] P. W. Anderson, G. Yuval, D. R. Hammann, Exact results in the Kondo problem. II. Scaling theory, qualitatively +correct solution, and some new results on one-dimensional classical statistical models, Phys. Rev. B 1 (1970) +4464. +[321] N. Read, D. M. Newns, On the solution of the Coqblin-Schreiffer Hamiltonian by the large-N expansion technique, +Journal of Physics C: Solid State Physics 16 (1983) 3273. +[322] P. Fendley, A. W. W. Ludwig, H. Saleur, Exact Conductance through Point Contacts in the ν = 1/3 Fractional +Quantum Hall Effect, Phys. Rev. Lett. 74 (1995) 3005. +[323] J. B. Miller, I. P. Radu, D. Z. Zumb¨uhl, E. Levenson-Falk, M. A. Kastner, C. M. Marcus, L. N. Pfeiffer, K. W. +West, Fractional quantum Hall effect in a quantum point contact at filling fraction 5/2, Nature Physics 3 (2007) +561. +[324] I. Radu, J. B. Miller, C. M. Marcus, M. A. Kastner, L. N. Pfeiffer, K. W. Baldwin, K. W. West, Quasi-Particle +Properties from Tunneling in the ν = 5/2 Fractional Quantum Hall State, Science 320 (2008) 899. +[325] C. L. Kane, M. Fisher, Nonequilibrium noise and fractional charge in the quantum Hall effect, Phys. Rev. Lett. 72 +(1994) 724. +[326] C. Chamon, D. E. Freed, X. G. Wen, Nonequilibrium quantum noise in chiral Luttinger liquids, Phys. Rev. B 53 +(1996) 4033. +[327] P. Fendley, A. Ludwig, H. Saleur, Exact Nonequilibrium dc Shot Noise in Luttinger Liquids and Fractional Quan- +tum Hall Devices, Phys. Rev. Lett. 75 (1995) 2196. +[328] M. Dolev, M. Heiblum, V. Umansky, A. Stern, D. Mahalu, Observation of a quarter of an electron charge at the +ν = 5/2 quantum Hall state, Nature 452 (2008) 829. +[329] C. Chamon, D. E. Freed, S. A. Kivelson, S. L. Sondhi, X. G. Wen, Two point-contact interferometer for quantum +Hall systems, Physical Review B 55 (1997) 2331. +[330] F. E. Camino, W. Zhou, V. J. Goldman, Realization of a Laughlin quasiparticle interferometer: Observation of +fractional statistics, Physical Review B 72 (2005) 075342. +[331] B. I. Halperin, A. Stern, I. Neder, B. Rosenow, Theory of the Fabry-P´erot quantum Hall interferometer, Physical +Review B 83 (2011) 155440. +[332] P. Bonderson, K. Shtengel, J. K. Slingerland, Probing non-abelian statistics with quasiparticle interferometry, +Physical Review Letters 97 (2006) 016401. +[333] A. Stern, B. I. Halperin, Proposed Experiments to Probe the Non-Abelian ν = 5/2 Quantum Hall State, Phys. Rev. +Lett. 96 (2006) 016802. +[334] R. L. Willett, L. N. Pfeiffer, K. W. West, Measurement of filling factor 5/2 quasiparticle interference with obser- +vation of charge e/4 and e/2 period oscillations, Proceedings of the National Academy of Sciences 106 (2009) +123 + +8853–8858. +[335] R. L. Willett, C. Nayak, K. Shtengel, L. N. Pfeiffer, K. W. West, Magnetic-Field-Tuned Aharonov-Bohm Oscilla- +tions and Evidence for Non-Abelian Anyons at ν = 5/2, Phys. Rev. Lett. 111 (2013) 186401. +[336] R. L. Willett, K. Shtengel, C. Nayak, L. N. Pfeiffer, Y. J. Chung, M. L. Peabody, K. W. Baldwin, K. W. W. +West, Interference measurements of non-Abelian e/4 & Abelian e/2 quasiparticle braiding, (unpublished) (2019). +arXiv:\protect\vrulewidth0pt\protect\href{http://arxiv.org/abs/1905.10248}{arXiv:1905.10248}. +[337] D. R. Nelson, J. M. Kosterlitz, Universal Jump in the Superfluid Density of Two-Dimensional Superfluids, Phys. +Rev. Lett. 39 (1977) 1201–1205. +[338] P. R. Thomas, M. Stone, Nature of the phase transition in a non-linear O(2)3 model, Nuclear Physics B 144 (2) +(1978) 513–524. +[339] M. E. Peskin, Mandelstam-’t Hooft duality in abelian lattice models, Annals of Physics 113 (1978) 122–152. +[340] C. Dasgupta, B. I. Halperin, Phase Transition in a Lattice Model of Superconductivity, Phys. Rev. Lett. 47 (1981) +1556–1560. +[341] B. I. Halperin, T. C. Lubensky, S. keng Ma, First-Order Phase Transitions in Superconductors and Smectic-A +Liquid Crystals, Physical Review Letters 32 (1974) 292. +[342] M. Moshe, J. Zinn-Justin, Quantum field theory in the large N limit: a review, Physics Reports 385 (2003) 69–228. +[343] H. Goldman, E. Fradkin, Loop Models, Modular Invariance, and Three Dimensional Bosonization, Phys. Rev. B +97 (2018) 195112. +[344] E. Fradkin, F. A. Schaposnik, The Fermion-Boson Mapping in Three Dimensional Quantum Field Theory, Phys. +Lett. B 338 (1994) 253. +[345] C. P. Burgess, F. Quevedo, Bosonization as duality, Nucl. Phys. B 421 (1994) 373. +[346] A. Chan, T. L. Hughes, S. Ryu, E. Fradkin, Effective field theories for topological insulators by functional +bosonization, Phys. Rev. B 87 (2013) 085132. +[347] A. P. O. Chan, T. Kvorning, S. Ryu, E. Fradkin, Effective hydrodynamic field theory and condensation picture of +topological insulators, Phys. Rev. B 93 (2016) 155122. +[348] S. Giombi, S. Minwalla, S. Prakash, S. P. Trivedi, S. R. Wadia, X. Yin, Chern-Simons theory with vector fermion +matter, Eur. Phys. J. C 72 (2012) 1–65. +[349] S. Jain, S. Minwalla, S. Yokoyama, Chern Simons duality with a fundamental boson and fermion, Journal of High +Eenergy Physics 2013 (2013) 37. +[350] O. Aharony, G. Gur-Ari, R. Yacoby, d = 3 bosonic vector models coupled to Chern-Simons gauge theories, J. +High Energy Phys. 03 (2012) 37. +[351] O. Aharony, G. Gur-Ari, R. Yacoby, Correlation functions of large N Chern-Simons-Matter theories and bosoniza- +tion in three dimensions, J. High Energy Phys. 2012 (2012) 28. +[352] O. Aharony, Baryons, monopoles and dualities in Chern-Simons-matter theories, J. High Energy Phys. 02 (2016) +93. +[353] N. Seiberg, E. Witten, Gapped boundary phases of topological insulators via weak coupling, Progress of Theoret- +ical and Experimental Physics 2016 (2016) ptw083. +[354] A. Karch, D. Tong, Particle-Vortex Duality from 3D Bosonization, Phys. Rev. X 6 (2016) 031043. +[355] D. T. Son, Is the composite fermion a Dirac particle?, Phys. Rev. X 5 (2015) 031027. +[356] M. A. Metlitski, A. Vishwanath, Particle-vortex duality of two-dimensional dirac fermion from electric-magnetic +duality of three-dimensional topological insulators, Phys. Rev. B 93 (2016) 245151. +[357] C. Wang, T. Senthil, Dual Dirac liquid on the surface of the electron topological insulator, Phys. Rev. X 5 (2015) +041031. +[358] J.-Y. Chen, J. H. Son, C. Wang, S. Raghu, Exact Boson-Fermion Duality on a 3D Euclidean Lattice, Phys. Rev. +Lett. 120 (2018) 016602. +[359] D. F. Mross, J. Alicea, O. I. Motrunich, Explicit Derivation of Duality between a Free Dirac Cone and Quantum +Electrodynamics in (2 + 1) Dimensions, Physical Review Letters 117 (2016) 016802. +[360] D. Simmons-Duffin, The Conformal Bootstrap, in: J. Polchinski, P. Vieira, O. DeWolfe (Eds.), TASI 2015: Pro- +ceedings of the 2015 Theoretical Advanced Study Institute in Elementary Particle Physics, New Frontiers in Fields +and Strings, World Scientific, Singapore, 2017, pp. 1–74. +[361] H. Goldman, E. Fradkin, Dirac composite fermions and emergent reflection symmetry about even-denominator +filling fractions, Physical Review B 98 (2018) 165137. +[362] D. Bohm, D. Pines, A Collective Description of Electron Interactions: III. Coulomb Interactions in a Degenerate +Electron Gas, Phys. Rev. 92 (1953) 609–625. +[363] F. D. M. Haldane, Luttinger’s Theorem and Bosonization of the Fermi Surface, in: R. A. Broglia, J. R. Schrieffer +(Eds.), Proceedings of the International School of Physics “Enrico Fermi”, Course CXXI “Perspectives in Many- +Particle Physics”, International School of Physics “Enrico Fermi”, North-Holland, Amsterdam, The Netherlands, +1994, pp. 5–29, arXiv:cond-mat/0505529. +[364] A. Houghton, J. B. Marston, Bosonization and fermion liquids in dimensions greater than one, Phys. Rev. B 48 +(1993) 7790–7808. +[365] A. H. Castro Neto, E. Fradkin, Bosonization of the Low Energy Excitations of Fermi Liquids, Phys. Rev. Lett. 72 +(1994) 1393. +[366] A. H. Castro Neto, E. Fradkin, Bosonization of Fermi Liquids, Phys. Rev. B 49 (1994) 10877. +[367] A. Houghton, H.-J. Kwon, J. B. Marston, Multidimensional bosonization, Advances in Physics 49 (2000) 141– +228. +[368] L. V. Delacr´etaz, Y.-H. Du, U. Mehta, D. T. Son, Nonlinear bosonization of Fermi surfaces: The method of +coadjoint orbits, Phys. Rev. Research 4 (2022) 033131. +[369] Z. D. Shi, H. Goldman, D. V. Else, T. Senthil, Gifts from anomalies: Exact results for Landau phase transitions in +metals, SciPost Phys. 13 (2022) 102. +[370] T. Senthil, A. Vishwanath, L. Balents, S. Sachdev, M. P. A. Fisher, Deconfined Quantum Critical Points, Science +303 (2004) 1490. +[371] A. L. Fitzpatrick, S. Kachru, J. Kaplan, S. Raghu, Non-Fermi-liquid fixed point in a Wilsonian theory of quantum +critical metals, Phys. Rev. B 88 (2013) 125116. +[372] M. A. Metlitski, D. F. Mross, S. Sachdev, T. Senthil, Cooper pairing in non-Fermi liquids, Phys. Rev. B 91 (2015) +115111. +124 + +[373] V. Oganesyan, S. A. Kivelson, E. Fradkin, Quantum theory of the nematic Fermi liquid, Physical Review B 64 +(2001) 195109. +[374] M. J. Lawler, D. G. Barci, V. Fern´andez, E. Fradkin, L. Oxman, Non-perturbative behavior of the quantum phase +transition to a nematic Fermi liquid, Physical Review B 73 (2006) 085101. +[375] M. J. Lawler, E. Fradkin, Local quantum criticality at the nematic quantum phase transition, Physical Review B +75 (2007) 0333304. +[376] S. Lederer, Y. Schattner, E. Berg, S. A. Kivelson, Enhancement of Superconductivity near a Nematic Quantum +Critical Point, Phys. Rev. Lett. 114 (2015) 097001. +[377] S. A. Kivelson, E. Fradkin, V. J. Emery, Electronic liquid-crystal phases of a doped Mott insulator, Nature 393 +(1998) 550–553. +[378] E. Fradkin, S. A. Kivelson, M. J. Lawler, J. P. E. andAndrew P. Mackenzie, Nematic Fermi Fluids in Condensed +Matter Physics, Annual Review of Condensed Matter Physics 1 (2010) 153–178. +[379] E. Fradkin, S. A. Kivelson, J. M. Tranquada, Colloquium: Theory of intertwined orders in high temperature +superconductors, Reviews of Modern Physics 87 (2015) 457–482. +[380] R. M. Fernandes, P. P. Orth, J. Schmalian, Intertwined Vestigial Order in Quantum Materials: Nematicity and +Beyond, Annual Review of Condensed Matter Physics 10 (2019) 133–154. +125 +