Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: mathdesign

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: CC BY 4.0
arXiv:2309.04358v3 [cond-mat.quant-gas] 14 Jan 2024

Beyond braid statistics: Constructing a lattice model for anyons with exchange statistics intrinsic to one dimension

Sebastian Nagies1,2,3, Botao Wang1,4, A.C. Knapp5, André Eckardt1\star and N.L.  Harshman6\dagger

1 Technische Universität Berlin, Institut für Theoretische Physik,

Hardenbergstr. 36, 10623 Berlin

2 Pitaevskii BEC Center and Department of Physics, University of Trento,

Via Sommarive 14, 38123 Trento, Italy

3 INFN-TIFPA, Trento Institute for Fundamental Physics and Applications, Trento, Italy

4 CENOLI, Université Libre de Bruxelles, CP 231, Campus Plaine, B-1050 Brussels, Belgium

5 University of Florida, Gainesville, FL 32610, USA

6 Physics Department, American University, Washington, DC 20016, USA

{}^{\star}start_FLOATSUPERSCRIPT ⋆ end_FLOATSUPERSCRIPT eckardt@tu-berlin.de, {}^{\dagger}start_FLOATSUPERSCRIPT † end_FLOATSUPERSCRIPT harshman@american.edu

Abstract

Anyons obeying fractional exchange statistics arise naturally in two dimensions: hard-core two-body constraints make the configuration space of particles not simply-connected. The braid group describes how topologically-inequivalent exchange paths can be associated to non-trivial geometric phases for abelian anyons. Braid-anyon exchange statistics can also be found in one dimension (1D), but this requires broken Galilean invariance to distinguish different ways for two anyons to exchange. However, recently it was shown that an alternative form of exchange statistics can occur in 1D because hard-core three-body constraints also make the configuration space not simply-connected. Instead of the braid group, the topology of exchange paths and their associated non-trivial geometric phases are described by the traid group. In this article we propose a first concrete model realizing this alternative form of anyonic exchange statistics. Starting from a bosonic lattice model that implements the desired geometric phases with number-dependent Peierls phases, we then define anyonic operators so that the kinetic energy term in the Hamiltonian becomes local and quadratic with respect to them. The ground-state of this traid-anyon-Hubbard model exhibits several indications of exchange statistics intermediate between bosons and fermions, as well as signs of emergent approximate Haldane exclusion statistics. The continuum limit results in a Galilean invariant Hamiltonian with eigenstates that correspond to previously constructed continuum wave functions for traid anyons. This provides not only an a-posteriori justification of our lattice model, but also shows that our construction serves as an intuitive approach to traid anyons, i.e. anyons intrinsic to 1D.

 

 

1 Introduction

How do wave functions transform when indistinguishable particles are exchanged? In three dimensions and higher, this question has two answers: symmetric under pairwise exchanges like bosons or antisymmetric under pairwise exchanges like fermions. In either case, the transformation does not depend on the path taken by the exchange, only on the permutation. However in lower dimensions, the quasiparticles that emerge in interacting systems can possess different forms of exchange statistics. The explanation is topological: unlike in three dimensions, particle interactions in one dimension (1D) and two dimensions (2D) create defects that make configuration space not simply connected. Different exchange paths for the same permutation can be topologically distinguished by how they wind around these defects. Anyons are particles with path-dependent exchange transformations, and they have statistics that lie intermediate between bosons and fermions.

Anyonic exchange statistics were first predicted for particles in 2D with hard-core two-body constraints [1, 2]. Such interactions create defects in configuration space and the braid group describes how topologically-distinct exchange paths can wind around these defects [3, 4, 5, 6]. Representations of the braid group determine how wave functions transform. The simplest abelian representations of the braid group give anyon wave functions with fractional exchange statistics described by a statistical angle θ[0,2π)𝜃02𝜋\theta\in[0,2\pi)italic_θ ∈ [ 0 , 2 italic_π ) that interpolates between bosons θ=0𝜃0\theta=0italic_θ = 0 and fermions θ=π𝜃𝜋\theta=\piitalic_θ = italic_π. Two-dimensional anyons obeying fractional statistics can be realized as quasiparticles of topologically ordered states of matter, such as fractional quantum Hall states [7, 8]. The anyonic braiding statistics with θ=2π/3𝜃2𝜋3\theta=2\pi/3italic_θ = 2 italic_π / 3 have been measured in the fractional quantum Hall effect through interference and scattering measurements [9, 10].

In 1D, interactions are even more disruptive to the topology of configuration space; particles must pass through each other in order to exchange. Hard-core two-body interactions make exchange impossible in 1D, so implementations of fractional exchange statistics in 1D must find alternate methods to implement the required geometrical phases. The anyon-Hubbard model [11, 12, 13], which was recently implemented with ultracold atoms in optical lattices [14], uses number-dependent Peierls-phases to generate the required phases. These phases break spatial parity, time-reversal, and Galilean symmetry, and even in the continuum limit of the anyon-Hubbard model, these symmetries are not restored [15].

However, it was recently demonstrated that a different form of anyonic exchange statistics intrinsic to 1D is possible if the two-body hard-core relation is relaxed but a three-body hard-core constraint is enforced [16, 17]. Hard-core three-body constraints in one-dimension also make the configuration space not simply-connected, and this allows for multi-valued wave functions and non-trivial exchange phases. Such a system is characterized by the so-called traid group which, like the braid group, can be visualized by strand diagrams with equivalence relations 111The strand group TNsubscript𝑇𝑁T_{N}italic_T start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT was independently discovered by several groups of mathematicians and goes by various other names, including the doodle group, planar braid group, and twin group [18, 19, 20, 21, 22].. Similar to fractional statistics and the braid group, abelian traid statistics are also intermediate between bosons and fermions, but in a different manner. The traid group breaks the Yang-Baxter relations, so that pairwise exchanges of some neighboring particle pairs can be symmetric and antisymmetric for others, for instance in a staggered fashion with respect to the particle order (as is explained below and in Refs. [16, 17]).

So far traid-anyonic exchange statistics has been studied only for quantum states in the continuum, where they are described by multi-valued wave functions. In this paper, we construct the first lattice model for traid anyons. For this purpose, we start from a one-dimensional bosonic lattice Hamiltonian with non-local, number-dependent tunneling phases. The phases embody the exchange properties of abelian representations of the traid group in a way reminiscent of bosonic descriptions of two-dimensional anyons via flux tube attachments and magnetic potentials. Using a non-local Jordan-Wigner-like gauge transformation, the bosonic model is then transformed to a local Hubbard-type Hamiltonian of anyonic particles obeying non-trivial, non-local commutation relations. The continuum limit of the bosonic representation of our model is found to give rise to solutions that correspond to the previously constructed traid anyonic wave functions, justifying our construction. We also find that, unlike the anyon-Hubbard model, the continuum limit of our traid-anyon-Hubbard model is Galilean invariant.

The motivation for our work is threefold: (i) We want to demonstrate that traid exchange statistics can be realized in a concrete model system. (ii) By directly engineering the required geometric phases associated with particle exchange, we want to provide an intuitive approach to the physics of traid anyons. (iii) Finally, we hope that our model might be a first step towards an experimental implementation of traid anyons, using techniques of quantum engineering in artificial quantum systems. While we do not propose a specific implementation for our model, we do identify the necessary structure of the number-dependent tunneling phases that an implementation would require. Due to the non-local many-body interactions involved, a prospective implementation will presumably be more complicated than the recently-realized anyon-Hubbard model [14]. Apart from (i-iii), we also make the interesting observation that our model, which by construction describes particles with non-trivial traid-anyonic exchange statistics, also shows indications of approximate fractional Haldane-type exclusion statistics [23, 24, 25].

The paper is organized as follows: In Sec. 2, we review how non-trivial exchange statistics can be described by strand diagrams governed by rules that emerge from the topology of configuration space. In doing so, we compare abelian braid and traid anyons to particles obeying standard (bosonic or fermionic) statistics as well as to each other. In Sec. 3, we then construct the traid-anyon-Hubbard model by starting from a bosonic model with number-dependent Peierls phases. Numerical results presented in Sec. 4 then allow us to study the ground-state properties of the model. We find generalized Friedel oscillations as well as other indications for fractional exclusion statistics in both the dependence of the chemical potential on the total particle number and the occupations of the natural orbits. In Sec. 5 we finally take the continuum limit and show that in this limit the solutions of our model coincide with the previously predicted wave functions for traid anyons. The concluding section puts our work into a broader context and points towards future research.

2 Exchange statistics in continuum models

As a prelude to the lattice models in the next section, in this section we consider exchange statistics in continuum systems in first quantization. We first review the Symmetrization Postulate and how the symmetric group describes permutations of fermions and bosons. Then we show how topological defects created by hard-core interactions in low dimensions can break the relations of the symmetric group, leading to the braid group and traid group. The abelian representations of these groups provide multi-valued anyonic wave functions that, unlike the symmetric group, have path-dependent exchange transformations. These results can be rigorously derived using the so-called intrinsic approach to topological exchange statistics; see Appendix A for an overview and some references.

2.1 Standard statistics

Bosons and fermions, whether single-component or multi-component, have standard exchange statistics. Unlike anyons, there is no distinction required between particle permutations and particle exchanges; all exchange paths leading to the same permutation are equivalent. For standard exchange statistics, the Symmetrization Postulate is sufficient to describe wave functions of bosons and fermions. To implement the Symmetrization Postulate, bosons or fermions are given arbitrary particle labels, temporarily breaking indistinguishability. Then the artificial labels are made unobservable by symmetrizing or antisymmetrizing the wave function over these labels, effectively restoring indistinguishibility.

In more detail, consider N𝑁Nitalic_N particles with labels i=1,2,,N𝑖12𝑁i=1,2,\ldots,Nitalic_i = 1 , 2 , … , italic_N and positions xisubscript𝑥𝑖x_{i}\in{\mathcal{M}}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∈ caligraphic_M summarized in the vector 𝐱=(x1,,xN)𝐱subscript𝑥1subscript𝑥𝑁{\bf x}=(x_{1},\ldots,x_{N})bold_x = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ). When there are no points excluded by hard-core interactions then the configuration space is 𝒳=N𝒳superscript𝑁{\mathcal{X}}={\mathcal{M}}^{N}caligraphic_X = caligraphic_M start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT, otherwise 𝒳N𝒳superscript𝑁{\mathcal{X}}\subset{\mathcal{M}}^{N}caligraphic_X ⊂ caligraphic_M start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT. The permutations p={p1p2pN}𝑝subscript𝑝1subscript𝑝2subscript𝑝𝑁p=\{p_{1}p_{2}\cdots p_{N}\}italic_p = { italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⋯ italic_p start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT } of these labels form the symmetric group SNsubscript𝑆𝑁S_{N}italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT and correspond to coordinate transformations on 𝒳𝒳{\mathcal{X}}caligraphic_X that map the point 𝐱=(x1,x2,,xN)𝐱subscript𝑥1subscript𝑥2subscript𝑥𝑁{\bf x}=(x_{1},x_{2},\ldots,x_{N})bold_x = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) to the point 𝐱=p(𝐱)=(xp1,xp2,,xpN)superscript𝐱𝑝𝐱subscript𝑥subscript𝑝1subscript𝑥subscript𝑝2subscript𝑥subscript𝑝𝑁{\bf x^{\prime}}=p({\bf x})=(x_{p_{1}},x_{p_{2}},\ldots,x_{p_{N}})bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_p ( bold_x ) = ( italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_POSTSUBSCRIPT ). Any permutation pSN𝑝subscript𝑆𝑁p\in S_{N}italic_p ∈ italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT can be expressed as a sequence of pairwise exchanges sisubscript𝑠𝑖s_{i}italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT with i=1,2,N1𝑖12𝑁1i=1,2\ldots,N-1italic_i = 1 , 2 … , italic_N - 1, each of which swaps the particle labelled i𝑖iitalic_i with the particle labelled (i+1)𝑖1(i+1)( italic_i + 1 ). The transpositions sisubscript𝑠𝑖s_{i}italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT form a generating set for the symmetric group and obey the relations

Self-inverse:si2=1,Self-inverse:superscriptsubscript𝑠𝑖21\displaystyle\mbox{Self-inverse:}\ s_{i}^{2}=1,Self-inverse: italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 , (1a)
Yang-Baxter:sisi+1si=si+1sisi+1,Yang-Baxter:subscript𝑠𝑖subscript𝑠𝑖1subscript𝑠𝑖subscript𝑠𝑖1subscript𝑠𝑖subscript𝑠𝑖1\displaystyle\mbox{Yang-Baxter:}\ s_{i}s_{i+1}s_{i}=s_{i+1}s_{i}s_{i+1},Yang-Baxter: italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT , (1b)
Locality:sisj=sjsiwhen|ij|>1.Locality:subscript𝑠𝑖subscript𝑠𝑗subscript𝑠𝑗subscript𝑠𝑖when𝑖𝑗1\displaystyle\mbox{Locality:}\ s_{i}s_{j}=s_{j}s_{i}\ \mbox{when}\ |i-j|>1.Locality: italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT when | italic_i - italic_j | > 1 . (1c)

A way to visualize these permutations is in terms of strand diagrams as shown in Fig. 1. Instead of a permutation of labels or a coordinate transformation on 𝒳𝒳{\mathcal{X}}caligraphic_X, the strand diagrams depict particle exchanges as N𝑁Nitalic_N continuous paths. The horizontal direction represents space and the vertical direction represents time (or a control parameter) increasing from bottom to top. The rules (2.1) can be understood intuitively as the equivalences achievable by continuously transforming the strands (moving and stretching in space, while maintaining forward progress in time) and allowing strands to cross and overlap arbitrarily. For the symmetric group, every strand diagram leading to the same permutation is equivalent under the rules (2.1); this is not true for braid and traid strand diagrams below.

Refer to caption
Figure 1: Examples of strand diagrams depicting equivalency relations (2.1). These are read from bottom to top, with particle 1111 furthest left. (a) self-inverse s12=1superscriptsubscript𝑠121s_{1}^{2}=1italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1; (b) Yang-Baxter s1s2s1=s2s1s2subscript𝑠1subscript𝑠2subscript𝑠1subscript𝑠2subscript𝑠1subscript𝑠2s_{1}s_{2}s_{1}=s_{2}s_{1}s_{2}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT; (c) locality s1s3=s3s1subscript𝑠1subscript𝑠3subscript𝑠3subscript𝑠1s_{1}s_{3}=s_{3}s_{1}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.

Applying the structural rules (2.1) to first-quantized wave functions for N𝑁Nitalic_N indistinguishable particles gives only two possibilities: bosons and fermions. To see this, consider the single-component (e.g., spinless or spin polarized) wave function ψ(𝐱)𝜓𝐱\psi({\bf x})italic_ψ ( bold_x ) defined on the configuration space 𝒳𝒳{\mathcal{X}}caligraphic_X. The probability |ψ(𝐱)|2superscript𝜓𝐱2|\psi({\bf x})|^{2}| italic_ψ ( bold_x ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to find particles at the positions 𝐱𝐱{\bf x}bold_x must be independent of the particle labels, so under the permutation p𝑝pitalic_p of particle labels the wave function can only change by a phase factor ηpsubscript𝜂𝑝\eta_{p}italic_η start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. Let the operator U^^𝑈\hat{U}over^ start_ARG italic_U end_ARG be a representation of SNsubscript𝑆𝑁S_{N}italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT on the Hilbert space of wave functions on 𝒳𝒳{\mathcal{X}}caligraphic_X defined by

U^pψ(𝐱)=ψ(p(𝐱))=ηpψ(𝐱).subscript^𝑈𝑝𝜓𝐱𝜓𝑝𝐱subscript𝜂𝑝𝜓𝐱\hat{U}_{p}\psi({\bf x})=\psi(p({\bf x}))=\eta_{p}\psi({\bf x}).over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_ψ ( bold_x ) = italic_ψ ( italic_p ( bold_x ) ) = italic_η start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_ψ ( bold_x ) . (2)

The phase factors ηpsubscript𝜂𝑝\eta_{p}italic_η start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT provide an abelian representation of the symmetric group, so they must obey the constraints related to Eqs. 2.1. Denoting σisubscript𝜎𝑖\sigma_{i}italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as the phase associated with the permutation sisubscript𝑠𝑖s_{i}italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, then an exchange of adjacent pairs gives U^siψ(𝐱)=σiψ(𝐱)subscript^𝑈subscript𝑠𝑖𝜓𝐱subscript𝜎𝑖𝜓𝐱\hat{U}_{s_{i}}\psi({\bf x})=\sigma_{i}\psi({\bf x})over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ψ ( bold_x ) = italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ψ ( bold_x ). From Eq. (1a), we have that ψ(si2(𝐱))𝜓superscriptsubscript𝑠𝑖2𝐱\psi(s_{i}^{2}({\bf x}))italic_ψ ( italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_x ) ) is equal to both σi2ψ(𝐱)superscriptsubscript𝜎𝑖2𝜓𝐱\sigma_{i}^{2}\psi({\bf x})italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ ( bold_x ) and ψ(𝐱)𝜓𝐱\psi({\bf x})italic_ψ ( bold_x ), implying that σi=±1subscript𝜎𝑖plus-or-minus1\sigma_{i}=\pm 1italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ± 1. In turn, from the Yang-Baxter relation (1b), we obtain ψ(sisi+1si(𝐱))=σi2σi+1ψ(𝐱)𝜓subscript𝑠𝑖subscript𝑠𝑖1subscript𝑠𝑖𝐱superscriptsubscript𝜎𝑖2subscript𝜎𝑖1𝜓𝐱\psi(s_{i}s_{i+1}s_{i}({\bf x}))=\sigma_{i}^{2}\sigma_{i+1}\psi({\bf x})italic_ψ ( italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_x ) ) = italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT italic_ψ ( bold_x ) is equal to ψ(si+1sisi+1(𝐱))=𝜓subscript𝑠𝑖1subscript𝑠𝑖subscript𝑠𝑖1𝐱absent\psi(s_{i+1}s_{i}s_{i+1}({\bf x}))=italic_ψ ( italic_s start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT ( bold_x ) ) = σiσi+12ψ(𝐱)subscript𝜎𝑖superscriptsubscript𝜎𝑖12𝜓𝐱\sigma_{i}\sigma_{i+1}^{2}\psi({\bf x})italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ ( bold_x ), implying that σi=σi+1subscript𝜎𝑖subscript𝜎𝑖1\sigma_{i}=\sigma_{i+1}italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT for all i𝑖iitalic_i. For abelian representations, locality (1c) provides no constraints on the phases. All together, we find that all phases are equal:

σi=σ,withσ=+1orσ=1,formulae-sequencesubscript𝜎𝑖𝜎withformulae-sequence𝜎1or𝜎1\sigma_{i}=\sigma,\quad\text{with}\quad\sigma=+1\quad\text{or}\quad\sigma=-1,italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_σ , with italic_σ = + 1 or italic_σ = - 1 , (3)

where the two possible choices correspond to bosons and fermions, respectively 222Note that a similar argument also holds for multi-component wave functions but requires non-abelian representations of the symmetric group. For a detailed exposition on the Symmetrization Postulate in 1D, see [26, 27]..

2.2 Braid anyons

In 2D with hard-core two-body interactions, the equivalence between permutations of particle labels and particle exchanges no longer holds. Instead of the symmetric group SNsubscript𝑆𝑁S_{N}italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT, particle exchanges for N𝑁Nitalic_N indistinguishable particles are described by the braid group BNsubscript𝐵𝑁B_{N}italic_B start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT [1, 2, 4, 5]. Because particles cannot coincide, the configuration space is not simply-connected and multi-valued wave functions are allowed; see Appendix B for more details on how the braid group derives from topological exchange statistics.

Refer to caption
Figure 2: Strand diagrams depicting self-inverse relations for the symmetric group and traid group and the broken self-inverse relations for the braid group. Subfigure (a) and (d) depicts the self-inverse relations for sisubscript𝑠𝑖s_{i}italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, the pairwise exchange of particle i𝑖iitalic_i and i+1𝑖1i+1italic_i + 1. Subfigures (b) and (c) depict the broken self-inverse relation for the pairwise exchanges bibi1subscript𝑏𝑖superscriptsubscript𝑏𝑖1b_{i}\neq b_{i}^{-1}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≠ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. The two-body hard-core constraint means that strands cannot overlap, forcing distinction between pairwise exchanges that are right-handed bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and left-handed bi1superscriptsubscript𝑏𝑖1b_{i}^{-1}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. If the particles live in three or more dimensions, these strands represent particle exchanges that could be continuously deformed into one another. In just one dimension, paths cannot go over or under, and so pairwise exchanges are topologically forbidden if there are hard-core two-body interactions. As we argue below, the anyon-Hubbard model allows to effectively define braid anyons also without a hard-core constraint in 1D.

The key structural features of the braid group can be understood by considering strand diagrams. Unlike the symmetric group, when two paths in the diagram cross, the particles cannot coincide and so we must indicate which particle is passing in front. As a result, the exchange of two particles is no longer self-inverse; the braids bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and bi1superscriptsubscript𝑏𝑖1b_{i}^{-1}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT are the distinct over- and under-braided exchanges of the particles labeled i𝑖iitalic_i and i+1𝑖1i+1italic_i + 1; see Fig. 2(b) and (c). With this distinction, the diagrams of Fig. 1 have to be modified as shown in Fig. 3. However, both the Yang-Baxter and locality relations are preserved because, unlike the self-inverse relation, continuous deformations of the strands described by those relations are not disrupted by the two-body hard-core constraints. Note that in three spatial dimensions also for hard-core interactions no distinction between over and under can be made since both diagrams can be transformed into each other using the third dimension. In one dimension, in turn, two-body hardcore interactions prevent any particle exchange. Thus, in the continuum, braid anyons are naturally expected to occur only in 2D. However, we show below how the discrete configuration space of a lattice allows for the definition of braid anyons also in 1D.

Refer to caption
Figure 3: Strand diagrams depicting the braid group. Subfigure (a) depicts the consequence b121superscriptsubscript𝑏121b_{1}^{2}\neq 1italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≠ 1 of the broken self-inverse relation bibi1subscript𝑏𝑖superscriptsubscript𝑏𝑖1b_{i}\neq b_{i}^{-1}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≠ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Subfigures (b) depicts the Yang-Baxter b1b2b1=b2b1b2subscript𝑏1subscript𝑏2subscript𝑏1subscript𝑏2subscript𝑏1subscript𝑏2b_{1}b_{2}b_{1}=b_{2}b_{1}b_{2}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT relation and subfigure (c) depicts the locality relation b1b3=b3b1subscript𝑏1subscript𝑏3subscript𝑏3subscript𝑏1b_{1}b_{3}=b_{3}b_{1}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_b start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, both unchanged from (1).

Every particle exchange γBN𝛾subscript𝐵𝑁\gamma\in B_{N}italic_γ ∈ italic_B start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT can be represented as the product of bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and bi1superscriptsubscript𝑏𝑖1b_{i}^{-1}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT’s. These generators obey the relations

Braid inequivalence:bi21,Braid inequivalence:superscriptsubscript𝑏𝑖21\displaystyle\mbox{Braid inequivalence:}\ b_{i}^{2}\neq 1,Braid inequivalence: italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≠ 1 , (4a)
Yang-Baxter:bibi+1bi=bi+1bibi+1,Yang-Baxter:subscript𝑏𝑖subscript𝑏𝑖1subscript𝑏𝑖subscript𝑏𝑖1subscript𝑏𝑖subscript𝑏𝑖1\displaystyle\mbox{Yang-Baxter:}\ b_{i}b_{i+1}b_{i}=b_{i+1}b_{i}b_{i+1},Yang-Baxter: italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_b start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT , (4b)
Locality:bibj=bjbiwhen|ij|>1.Locality:subscript𝑏𝑖subscript𝑏𝑗subscript𝑏𝑗subscript𝑏𝑖when𝑖𝑗1\displaystyle\mbox{Locality:}\ b_{i}b_{j}=b_{j}b_{i}\ \mbox{when}\ |i-j|>1.Locality: italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT when | italic_i - italic_j | > 1 . (4c)

Note that unlike the symmetric group, the same particle permutation can be enacted by multiple inequivalent exchange paths γBN𝛾subscript𝐵𝑁\gamma\in B_{N}italic_γ ∈ italic_B start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT. For example, the braids b1subscript𝑏1b_{1}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, b11superscriptsubscript𝑏11b_{1}^{-1}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, b13superscriptsubscript𝑏13b_{1}^{3}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, b13superscriptsubscript𝑏13b_{1}^{-3}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, etc., all enact the same particle permutation of the first and second particles. The multi-valuedness of anyon wave functions on 𝒳𝒳{\mathcal{X}}caligraphic_X originates from the multiplicity of topologically distinct exchange paths that execute the same permutation. Unlike the Symmetrization Postulate, elements of the braid group do not correspond to coordinate transformations on 𝒳𝒳{\mathcal{X}}caligraphic_X. However, we can define an operator U^γsubscript^𝑈𝛾\hat{U}_{\gamma}over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT that represents a particle exchange γBN𝛾subscript𝐵𝑁\gamma\in B_{N}italic_γ ∈ italic_B start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT, and for an abelian representation, U^bisubscript^𝑈subscript𝑏𝑖\hat{U}_{b_{i}}over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT multiplies the wave function by a phase factor βisubscript𝛽𝑖\beta_{i}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT:

U^biψ(𝐱)=βiψ(𝐱).subscript^𝑈subscript𝑏𝑖𝜓𝐱subscript𝛽𝑖𝜓𝐱\hat{U}_{b_{i}}\psi({\bf x})=\beta_{i}\psi({\bf x}).over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ψ ( bold_x ) = italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ψ ( bold_x ) . (5)

Because the generators are no longer self-inverse, the phase βisubscript𝛽𝑖\beta_{i}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT can take any value. As a result of the Yang-Baxter relation, as before we have βi+1=βisubscript𝛽𝑖1subscript𝛽𝑖\beta_{i+1}=\beta_{i}italic_β start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT = italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and all the βisubscript𝛽𝑖\beta_{i}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are required to be equal. Therefore, we obtain the continuous family of possible choices for abelian representations of BNsubscript𝐵𝑁B_{N}italic_B start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT

βi=eiθwithθ[0,2π),formulae-sequencesubscript𝛽𝑖superscript𝑒𝑖𝜃with𝜃02𝜋\beta_{i}=e^{i\theta}\quad\text{with}\quad\theta\in[0,2\pi),italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT with italic_θ ∈ [ 0 , 2 italic_π ) , (6)

labeled by the exchange phase θ𝜃\thetaitalic_θ. The special cases θ=0𝜃0\theta=0italic_θ = 0 and θ=π𝜃𝜋\theta=\piitalic_θ = italic_π correspond to bosons and fermions, respectively, and lead to single-valued wave functions on 𝒳𝒳{\mathcal{X}}caligraphic_X. All other θ𝜃\thetaitalic_θ correspond to braid anyons with fractional exchange statistics and have wave functions that are multi-valued on 𝒳𝒳{\mathcal{X}}caligraphic_X [6].

2.3 Traid anyons

The equivalence between permutations of particle labels and particle exchanges also no longer holds when there are three-body hard-core interactions in 1D. The excluded points in configuration space are defects that result in topological exchange statistics described by the traid group TNsubscript𝑇𝑁T_{N}italic_T start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT [16, 17]. See App. C for how the traid group is derived from topological exchange statistics.

To understand the resulting exchange statistics, we can again modify the strand diagrams of Fig. 1, as shown in Fig. 4. Since we do not assume a two-body hard-core constraint, two strands can cross, just like in Fig. 1; there is no “over” or “under” distinction intrinsic to 1D, unless we encode this information in additional degrees of freedom like relative velocity. We can immediately verify that the exchange of particles is self-inverse and local. However, the Yang-Baxter relation is broken; there is no continuous deformation of the strand diagram on the left-hand side of 1(b) to that on the right-hand side of 1(b). Such a deformation would necessarily involve crossing of three strands in one point, violating the three-body hard-core constraint.

Refer to caption
Figure 4: Strand diagrams depicting the traid group. Subfigures (a) and (c) depict the self-inverse t12=1superscriptsubscript𝑡121t_{1}^{2}=1italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 and locality relations t1t3=t3t1subscript𝑡1subscript𝑡3subscript𝑡3subscript𝑡1t_{1}t_{3}=t_{3}t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_t start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, both unchanged from (1). Subfigure (b) depicts the broken Yang-Baxter relation t1t2t1t2t1t2subscript𝑡1subscript𝑡2subscript𝑡1subscript𝑡2subscript𝑡1subscript𝑡2t_{1}t_{2}t_{1}\neq t_{2}t_{1}t_{2}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≠ italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. The absence of a two-body hard-core constraint means that two strands can overlap, but the three-body hard-core constraint implies that three strands cannot. Therefore, a continuous deformation from the strand diagram on the left to the strand diagram on the right is impossible, and these two exchange paths are topologically distinguishable and may acquire different phases, despite leading to the same permutation.

As before, we can define generators tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for the exchange of the i𝑖iitalic_ith and (i+1)𝑖1(i+1)( italic_i + 1 )th particles that obey relations that we can read off Fig. 4:

Self-inverse:ti2=1,Self-inverse:superscriptsubscript𝑡𝑖21\displaystyle\mbox{Self-inverse:}\ t_{i}^{2}=1,Self-inverse: italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 , (7a)
Traid inequivalence:titi+1titi+1titi+1,Traid inequivalence:subscript𝑡𝑖subscript𝑡𝑖1subscript𝑡𝑖subscript𝑡𝑖1subscript𝑡𝑖subscript𝑡𝑖1\displaystyle\mbox{Traid inequivalence:}\ t_{i}t_{i+1}t_{i}\neq t_{i+1}t_{i}t_% {i+1},Traid inequivalence: italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≠ italic_t start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT , (7b)
Locality:titj=tjtiwhen|ij|>1.Locality:subscript𝑡𝑖subscript𝑡𝑗subscript𝑡𝑗subscript𝑡𝑖when𝑖𝑗1\displaystyle\mbox{Locality:}\ t_{i}t_{j}=t_{j}t_{i}\ \mbox{when}\ |i-j|>1.Locality: italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT when | italic_i - italic_j | > 1 . (7c)

With these relations, the N1𝑁1N-1italic_N - 1 generators tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT define the traid group TNsubscript𝑇𝑁T_{N}italic_T start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT for N𝑁Nitalic_N strands. Note that while for the symmetric group the labels i𝑖iitalic_i were an arbitrary assignment, and for the braid group the labels were based on an arbitrary positioning of the particles, in 1D there is a natural way to assign labels based on their ordering on the line; i.e. the particle furthest to the left (or equivalently, right) is the particle labeled i=1𝑖1i=1italic_i = 1, then the next particle in order is labeled i=2𝑖2i=2italic_i = 2, and so on. This ordering scheme is invariant under homeomorphisms of the line, and ordinality (so defined) is a topological property intrinsic to 1D systems. Therefore, unlike the pairwise exchange generators of SNsubscript𝑆𝑁S_{N}italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT or BNsubscript𝐵𝑁B_{N}italic_B start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT, the generator tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT exchanges neighboring particles in real space, not just in label space. This topological feature of ordinality is intimately connected with the breaking of the Yang-Baxter relation (7b).

Irreducible abelian representations of the traid group are specified by the phases τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT picked up during the exchange of neighboring particles

U^tiψ(𝐱)=τiψ(𝐱).subscript^𝑈subscript𝑡𝑖𝜓𝐱subscript𝜏𝑖𝜓𝐱\hat{U}_{t_{i}}\psi({\bf x})=\tau_{i}\psi({\bf x}).over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ψ ( bold_x ) = italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ψ ( bold_x ) . (8)

where the operator U^tisubscript^𝑈subscript𝑡𝑖\hat{U}_{t_{i}}over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT represents the generator tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The self-inverse relation (7a) ensures that, like for standard SNsubscript𝑆𝑁S_{N}italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT statistics, τi=±1subscript𝜏𝑖plus-or-minus1\tau_{i}=\pm 1italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ± 1. However, unlike both standard and braid statistics, where the Yang-Baxter relation (7b) enforces equality of the phases, for the traid group the phase factors τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are independent (generally, τiτjsubscript𝜏𝑖subscript𝜏𝑗\tau_{i}\neq\tau_{j}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≠ italic_τ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT). Thus, we have

τi=+1orτi=1,formulae-sequencesubscript𝜏𝑖1orsubscript𝜏𝑖1\tau_{i}=+1\quad\text{or}\quad\tau_{i}=-1,italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = + 1 or italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = - 1 , (9)

leading to 2N1superscript2𝑁12^{N-1}2 start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT abelian representations of the traid group, corresponding to the N1𝑁1N-1italic_N - 1 choices of signs τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Like braid anyons, traid anyons contain bosons and fermions as special cases, here corresponding to all τi=+1subscript𝜏𝑖1\tau_{i}=+1italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = + 1 or all τi=1subscript𝜏𝑖1\tau_{i}=-1italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = - 1, respectively. Unlike braid anyons, traid anyon representations do not continuously interpolate between these two extreme cases, but instead comprise a discrete set with non-standard exchange statistics.

To better understand traid anyon representations, consider the simplest non-trivial case of N=3𝑁3N=3italic_N = 3 particles. Besides bosons (τ1,τ2)=(+1,+1)(++)(\tau_{1},\tau_{2})=(+1,+1)\equiv(++)( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = ( + 1 , + 1 ) ≡ ( + + ) and fermions ()(--)( - - ), there are two additional abelian representations (+)(+-)( + - ) and (+)(-+)( - + ). For the representation (+)(+-)( + - ), the first pair of particles exchanges symmetrically like bosons and the second pair exchanges antisymmetrically like fermions. For either of these cases, the products τ1τ2τ1subscript𝜏1subscript𝜏2subscript𝜏1\tau_{1}\tau_{2}\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and τ2τ1τ2subscript𝜏2subscript𝜏1subscript𝜏2\tau_{2}\tau_{1}\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT have opposite signs, showing that the same permutation of the first and third particle can be accomplished through topologically distinct exchange paths.

3 Lattice model for traid anyons

The path-dependence of exchange statistics in one-dimensional tight-binding lattice models is in some ways more intuitive and easier to visualize than the continuum case. The configuration space becomes a discrete Fock space and exchange paths become a series of discrete hops with dynamical phases. Exchange paths can be understood as loops in the discrete configuration space given by the Fock states, and the total dynamical phase can be visualized as a flux through this loop.

By engineering these phases and fluxes, bosonic lattice models can implement non-standard exchange statistics with density-dependent tunneling phases. In the first subsection we first recapitulate how this works for the anyon-Hubbard model, which was introduced to simulate fractional exchange statistics with bosons on a one-dimensional lattice [11]. In the next subsection, we reverse engineer the argument that led to the anyon-Hubbard model to create a Hubbard model with abelian traid-anyon statistics.

3.1 Anyon-Hubbard model with braid statistics

To introduce the anyon-Hubbard model, we follow Ref. [11] and begin from the deformed commutation relations hypothesized for fractional exchange statistics:

ajakakeiθsgn(jk)aj=0,subscript𝑎𝑗subscript𝑎𝑘subscript𝑎𝑘superscript𝑒𝑖𝜃sgn𝑗𝑘subscript𝑎𝑗0\displaystyle a_{j}a_{k}-a_{k}e^{i\theta\text{sgn}(j-k)}a_{j}=0,italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_a start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_θ sgn ( italic_j - italic_k ) end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 0 , (10)
ajakakeiθsgn(jk)aj=δjk,subscript𝑎𝑗superscriptsubscript𝑎𝑘superscriptsubscript𝑎𝑘superscript𝑒𝑖𝜃sgn𝑗𝑘subscript𝑎𝑗subscript𝛿𝑗𝑘\displaystyle a_{j}a_{k}^{\dagger}-a_{k}^{\dagger}e^{-i\theta\text{sgn}(j-k)}a% _{j}=\delta_{jk},italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT - italic_a start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_θ sgn ( italic_j - italic_k ) end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT ,

where ajsubscript𝑎𝑗a_{j}italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the annihilation operator for an anyon obeying fractional exchange statistics described by exchange phase θ𝜃\thetaitalic_θ. In terms of these operators, the anyon-Hubbard Hamiltonian is then

H=Jj(aj+1aj+h.c.)+U2jnj(nj1),𝐻𝐽subscript𝑗subscriptsuperscript𝑎𝑗1subscript𝑎𝑗h.c.𝑈2subscript𝑗subscript𝑛𝑗subscript𝑛𝑗1H=-J\sum_{j}\left(a^{\dagger}_{j+1}a_{j}+\textrm{h.c.}\right)+\frac{U}{2}\sum_% {j}n_{j}(n_{j}-1),italic_H = - italic_J ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + h.c. ) + divide start_ARG italic_U end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - 1 ) , (11)

where nj=ajajsubscript𝑛𝑗superscriptsubscript𝑎𝑗subscript𝑎𝑗n_{j}=a_{j}^{\dagger}a_{j}italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the number operator on site j𝑗jitalic_j, J𝐽Jitalic_J is the tunneling strength, and U𝑈Uitalic_U is the two-body interaction strength.

The anyonic creation and annhihilation operators with commutation relations (10) can be expressed in terms of bosonic ones bjsubscript𝑏𝑗b_{j}italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT via a generalized gauge transformation called the fractional Jordan-Wigner transformation:

aj=eiθkjnkbj,subscript𝑎𝑗superscript𝑒𝑖𝜃subscript𝑘𝑗subscript𝑛𝑘subscript𝑏𝑗a_{j}=e^{i\theta\sum_{k\leq j}n_{k}}b_{j},italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT italic_i italic_θ ∑ start_POSTSUBSCRIPT italic_k ≤ italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (12)

where also nj=bjbjsubscript𝑛𝑗superscriptsubscript𝑏𝑗subscript𝑏𝑗n_{j}=b_{j}^{\dagger}b_{j}italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. This density-dependent gauge transformation (12) is non-local because the phase for the operator ajsubscript𝑎𝑗a_{j}italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT depends on all occupation numbers to the left of site j𝑗jitalic_j. Although the fractional Jordan-Wigner gauge transformation is non-local, the deformed commutation relations (10) remain local.

Refer to caption
Figure 5: (a) Sketch of density-dependent hopping processes in the anyon-Hubbard model (Eq. 13). (b) Four configurations of two particles on three lattice sites. The parity-breaking density-dependent hopping phase means that the transition from the leftmost configuration to the rightmost configuration that takes place without coincidence accumulates a different phase than the lower path. Mapping this back to Fig. 3a, a clockwise loop returning to the same configuration corresponds to the braid generator bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and a counterclockwise to bi1superscriptsubscript𝑏𝑖1b_{i}^{-1}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT.

Expressed in terms of boson operators, the anyon-Hubbard Hamiltonian (11) becomes

H=Jj(bj+1eiθnj+1bj+h.c.)+U2jnj(nj1).𝐻𝐽subscript𝑗subscriptsuperscript𝑏𝑗1superscript𝑒𝑖𝜃subscript𝑛𝑗1subscript𝑏𝑗h.c.𝑈2subscript𝑗subscript𝑛𝑗subscript𝑛𝑗1H\!=\!-J\sum_{j}\left(b^{\dagger}_{j+1}e^{-i\theta n_{j+1}}b_{j}\!+\!\textrm{h% .c.}\!\right)+\frac{U}{2}\!\sum_{j}\!n_{j}(n_{j}-1).italic_H = - italic_J ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_θ italic_n start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + h.c. ) + divide start_ARG italic_U end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - 1 ) . (13)

In the bosonic form, there are operator-valued, occupation-number-dependent Peierls phases attached to the tunneling matrix elements. These phases break parity, and they occur when a particle hops to the right onto an already occupied lattice site [Fig. 5(a)] or in the hermitian conjugate process when a particle hops to the left from a multiply occupied lattice site.

These Peierls phases can give rise to non-trivial geometric phases associated with closed paths in the many-body configuration space of indistinguishable particles, given by all Fock states with sharp site occupation numbers. Such geometric phases can be viewed as generalized magnetic fluxes. In Fig. 5(b), we show an example of a Fock-space loop associated with a nontrivial phase of θ𝜃\thetaitalic_θ (θ𝜃-\theta- italic_θ), when encircled in clockwise (anticlockwise) direction. Recently, the bosonic representation of the anyon-Hubbard model (13) was realized experimentally with ultracold atoms in an optical lattice, where the non-trivial geometric phases were probed directly via interference of the upper and the lower path of Fig. 5(b) during quantum walks [14].

The geometric phase resulting from the density-dependent tunneling can be directly related to the braiding of particles. In Fig. 5(b), we can associate paths that encircle the depicted loop, starting (say) from the leftmost configuration, with the exchange of the two particles. Encircling the loop in clockwise direction, so that the left particle tunnels rightwards twice, passing an occupied site, can be associated with bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT depicted in the strand diagram Fig. 2(b), where the left particle moves over the right one. In turn, encircling the loop in anticlockwise direction, so that the right particle tunnels leftwards twice, passing an occupied site, can be associated with bi1superscriptsubscript𝑏𝑖1b_{i}^{-1}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT depicted in the strand diagram Fig. 2(c), where the right particle moves over the left one. The corresponding phase factors are βi=β=eiθsubscript𝛽𝑖𝛽superscript𝑒𝑖𝜃\beta_{i}=\beta=e^{-i\theta}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_β = italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT and βi1=β*=eiθsuperscriptsubscript𝛽𝑖1superscript𝛽superscript𝑒𝑖𝜃\beta_{i}^{-1}=\beta^{*}=e^{i\theta}italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = italic_β start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT. In this sense, the anyon-Hubbard model describes braid anyons. Therefore, and in order to distinguish it from the traid-anyon-Hubbard model that we construct in the next subsection, we will also refer to it as the braid-anyon-Hubbard model in the following.

3.2 Hubbard model with traid exchange statistics

The goal of this section is to introduce a lattice model of interacting bosons that captures traid anyonic exchange statistics. For this purpose, we will reverse the strategy used for the braid-anyon-Hubbard model above to implement these extreme cases. Namely, we first construct a bosonic model with non-local number-dependent tunnelling phases that give rise to the desired exchange phases. Then we find the corresponding traid anyon field operators and their commutation relations by constructing a generalized Jordan-Wigner transformation, so that the kinetic energy term of the Hamiltonian becomes quadratic and local with respect to these new operators. We will first focus on the case of staggered abelian representations of the traid group, with alternating exchange phases given by (++)(+-+-\cdots)( + - + - ⋯ ) and (++)(-+-+\cdots)( - + - + ⋯ ). These are of specific interest, since they are furthest away both from the bosonic and (pseudo)fermionic cases, characterized by the homogeneous exchange phases (+++)(++\cdots+)( + + ⋯ + ) and ()(--\cdots-)( - - ⋯ - ), respectively. Subsequently, we will also construct a model for general abelian representations (τ1,τ2,,τN1)subscript𝜏1subscript𝜏2subscript𝜏𝑁1(\tau_{1},\tau_{2},\ldots,\tau_{N-1})( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_τ start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT ) with τi=+1subscript𝜏𝑖1\tau_{i}=+1italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = + 1 or 11-1- 1 of the traid group.

A model that realizes the two abelian representations of the traid group with alternating exchange phases is achieved by the following lattice model using bosons with number-dependent hopping phases,

H𝐻\displaystyle Hitalic_H =\displaystyle== JjL(bj+1eiπ(Nj+I)nj+1bj+h.c.)+U2jLnj(nj1).\displaystyle-J\sum_{j}^{L}\left(b^{\dagger}_{j+1}e^{i\pi\left(N_{j}+I\right)n% _{j+1}}b_{j}+\mathrm{h.c.}\right)+\frac{U}{2}\sum^{L}_{j}n_{j}(n_{j}-1).- italic_J ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT ( italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_π ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_I ) italic_n start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + roman_h . roman_c . ) + divide start_ARG italic_U end_ARG start_ARG 2 end_ARG ∑ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - 1 ) . (14)

In contrast to the braid-anyon-Hubbard model (13), the number-dependent hopping phase is now defined in terms of the operator

Njkjnk,subscript𝑁𝑗subscript𝑘𝑗subscript𝑛𝑘N_{j}\equiv\sum_{k\leq j}n_{k},italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≡ ∑ start_POSTSUBSCRIPT italic_k ≤ italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , (15)

that counts the number of particles to the left of (and including) lattice site j𝑗jitalic_j. The integer I{0,1}𝐼01I\in\{0,1\}italic_I ∈ { 0 , 1 } that appears in the phase determines which of the two possible abelian representations of the traid group with alternating exchange phases the model realizes: For I=0𝐼0I=0italic_I = 0 we have (+++)(+-+-+\cdots)( + - + - + ⋯ ), meaning that the two leftmost particles exchange like bosons, the second and third particles from the left exchange like fermions, and so on. For I=1𝐼1I=1italic_I = 1 on the other hand, the pattern (++)(-+-+\cdots)( - + - + ⋯ ) gets flipped, i.e. the two leftmost particles now exchange like fermions.

The symmetries of the lattice traid anyon model can be inferred from the bosonic form of the Hamiltonian (14). Time reversal invariance 𝒯H𝒯=H𝒯𝐻𝒯𝐻\mathcal{T}H\mathcal{T}=Hcaligraphic_T italic_H caligraphic_T = italic_H results by observing that exp(iπk)𝑖𝜋𝑘\exp(i\pi k)roman_exp ( italic_i italic_π italic_k ) =exp(iπk)absent𝑖𝜋𝑘=\exp(-i\pi k)= roman_exp ( - italic_i italic_π italic_k ) for any integer k𝑘kitalic_k, so the Peierls phase is invariant under complex conjugation, as are all other terms. Spatial inversion 𝒫𝒫\mathcal{P}caligraphic_P transforms j𝑗jitalic_j to Lj𝐿𝑗L-jitalic_L - italic_j and Njsubscript𝑁𝑗N_{j}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT to NNj𝑁subscript𝑁𝑗N-N_{j}italic_N - italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. After reindexing the sum, the only difference between 𝒫H𝒫𝒫𝐻𝒫\mathcal{P}H\mathcal{P}caligraphic_P italic_H caligraphic_P and H𝐻Hitalic_H that remains is an additional phase exp(iπN)𝑖𝜋𝑁\exp(i\pi N)roman_exp ( italic_i italic_π italic_N ) in the hopping terms. Therefore, for N𝑁Nitalic_N even, we have 𝒫H(I)𝒫=H(I)𝒫𝐻𝐼𝒫𝐻𝐼\mathcal{P}H(I)\mathcal{P}=H(I)caligraphic_P italic_H ( italic_I ) caligraphic_P = italic_H ( italic_I ) with the same I𝐼Iitalic_I whereas for N𝑁Nitalic_N odd we have 𝒫H(I=0)𝒫=H(I=1)𝒫𝐻𝐼0𝒫𝐻𝐼1\mathcal{P}H(I=0)\mathcal{P}=H(I=1)caligraphic_P italic_H ( italic_I = 0 ) caligraphic_P = italic_H ( italic_I = 1 ). In comparison, neither 𝒯𝒯\mathcal{T}caligraphic_T nor 𝒫𝒫\mathcal{P}caligraphic_P are symmetries of the braid-anyon-Hubbard model, although a gauge-transformed combination of them is [28].

Refer to caption
Figure 6: (a-c) Sketch of some possible hopping processes and associated phase factors in our lattice model for traid anyons (Eqs. 14). Depicted is a system of 4444 particles with an integer I=0𝐼0I=0italic_I = 0, corresponding to the abelian representation (+++-++ - +). (d-e) Two loops in configuration space for the N=3𝑁3N=3italic_N = 3 traid anyon model representation (+)(+-)( + - ), analgous to Fig. 5. Unlike the braid-anyon-Hubbard model, all hopping parameters are real, so there is no difference between clockwise and counterclockwise loops, i.e., these exchanges are self-inverse. In (d), the first and second particles undergo an exchange process. All hopping phases have the same sign and the product of all the phase factors is +11+1+ 1. In (e), the second and third particles undergo an exchange process. The sign of lower left hops is reverse, so the product of all the phase factors is 11-1- 1.

Figs. 6(a-c) depict these alternating hopping phases (++)(+-+)( + - + ) for the case of 4 particles and I=0𝐼0I=0italic_I = 0: When the leftmost particle hops to the right onto an already occupied lattice site [Fig. 6(a)], a phase factor of +11+1+ 1 gets picked up. Tunneling to the right onto an empty site always gives a trivial phase factor, and so if that particle were to then hop onto another empty site (not depicted), the two particles would be exchanged in order with an overall exchange phase factor of +11+1+ 1. Exchanging the third and fourth particle gives the same result, while a swap of the central two particles [Fig. 6(c)] results in an overall phase factor of 11-1- 1, i.e. the second and third particles exchange like fermions. Figs. 6 (d,e) show how these phases accumulate during particle exchanges. We can see that, like for the braid-anyon-Hubbard model, the number-dependent Peierls phases give rise to the desired exchange phases.

Note that our lattice model in Eq. 14 uses bosons and thus allows an arbitrary number of particles to occupy the same lattice site. However, a three-body hard-core constraint is essential for the existence of genuine traid anyons in the continuum. Indeed, if we consider hopping processes on the lattice involving three or more particles, the desired pattern of alternating exchange phases breaks down for our model. We can address this problem theoretically in two ways: we can add a three-body interaction term to the Hamiltonian (14) and take the limit of the coupling coefficient to infinity, or we can implement the three-body hard-core constraint algebraically by requiring bj3=0superscriptsubscript𝑏𝑗30b_{j}^{3}=0italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT = 0. For numerical convenience, we have chosen the latter approach by simply truncating the state space so that at most two particles can occupy the same site. Note that if we restrict ourselves to the low-energy, dilute limit, then three-body occupation of the same site becomes increasingly unlikely to occur and should be negligible for the calculation of most observables. Our numerical results in the next section show that this is indeed the case for the low-energy dilute limit. Further, in the following section on the continuum limit we show that three-body, hard-core interactions emerge naturally from the number-dependent tunneling phases for the alternating representations.

We now define annihilation and creation operators, tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and tjsuperscriptsubscript𝑡𝑗t_{j}^{\dagger}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, for the traid anyons. For that purpose, we define a generalized Jordan-Wigner transformation, so that the kinetic part of the Hamiltonian describing tunneling becomes local and quadratic with respect to these new operators:

H𝐻\displaystyle Hitalic_H =\displaystyle== JjL(tj+1tj+h.c.)+U2jLnj(nj1).\displaystyle-J\sum_{j}^{L}\left(t^{\dagger}_{j+1}t_{j}+\mathrm{h.c.}\right)+% \frac{U}{2}\sum^{L}_{j}n_{j}(n_{j}-1).- italic_J ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + roman_h . roman_c . ) + divide start_ARG italic_U end_ARG start_ARG 2 end_ARG ∑ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - 1 ) . (16)

For the staggered traid anyon model (14), this is achieved by:

tjsubscript𝑡𝑗\displaystyle t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT \displaystyle\equiv eiπMjbj,superscript𝑒𝑖𝜋subscript𝑀𝑗subscript𝑏𝑗\displaystyle e^{-i\pi M_{j}}b_{j},italic_e start_POSTSUPERSCRIPT - italic_i italic_π italic_M start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (17a)
Mjsubscript𝑀𝑗\displaystyle M_{j}italic_M start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT \displaystyle\equiv 12[kjnk(Njnk)]+INj,12delimited-[]subscript𝑘𝑗subscript𝑛𝑘subscript𝑁𝑗subscript𝑛𝑘𝐼subscript𝑁𝑗\displaystyle\frac{1}{2}\left[\sum_{k\leq j}n_{k}\left(N_{j}-n_{k}\right)% \right]+IN_{j},divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ ∑ start_POSTSUBSCRIPT italic_k ≤ italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_n start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ] + italic_I italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (17b)

where Mjsubscript𝑀𝑗M_{j}italic_M start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT depends on the Njsubscript𝑁𝑗N_{j}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT (15) and on the integer I𝐼Iitalic_I (determining which of the two representations with staggered exchange phases is realized in Eq. 14). Like the braid-anyon-Hubbard model, the transformation (3.2) leaves the number operator on site j𝑗jitalic_j invariant

tjtj=bjbj=nj.subscriptsuperscript𝑡𝑗subscript𝑡𝑗subscriptsuperscript𝑏𝑗subscript𝑏𝑗subscript𝑛𝑗t^{\dagger}_{j}t_{j}=b^{\dagger}_{j}b_{j}=n_{j}.italic_t start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT . (18)

The crucial difference is that when expressed in the bosonic form, the hopping phases of the traid anyon model (14) are non-local and depend on the notion of ordinality, whereas hopping phases for the braid-anyon-Hubbard model are local.

This non-local property is noticeable also in the commutation relations for the traid operators tjsubscript𝑡𝑗t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. From Eq. 3.2 and the bosonic commutation relations, we obtain:

tjtktkeiπsgn(jk)[Njk+I]tjsubscript𝑡𝑗subscript𝑡𝑘subscript𝑡𝑘superscript𝑒𝑖𝜋sgn𝑗𝑘delimited-[]subscript𝑁𝑗𝑘𝐼subscript𝑡𝑗\displaystyle t_{j}t_{k}-t_{k}e^{i\pi\text{sgn}(j-k)\left[N_{jk}+I\right]}t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_π sgn ( italic_j - italic_k ) [ italic_N start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT + italic_I ] end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT =\displaystyle== 0,0\displaystyle 0,0 , (19a)
tjtktkeiπsgn(jk)[Njk+I]tjsubscript𝑡𝑗superscriptsubscript𝑡𝑘superscriptsubscript𝑡𝑘superscript𝑒𝑖𝜋sgn𝑗𝑘delimited-[]subscript𝑁𝑗𝑘𝐼subscript𝑡𝑗\displaystyle t_{j}t_{k}^{\dagger}-t_{k}^{\dagger}e^{-i\pi\text{sgn}(j-k)\left% [N_{jk}+I\right]}t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT - italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_π sgn ( italic_j - italic_k ) [ italic_N start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT + italic_I ] end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT =\displaystyle== δjk,subscript𝛿𝑗𝑘\displaystyle\delta_{jk},italic_δ start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT , (19b)

where Njksubscript𝑁𝑗𝑘N_{jk}italic_N start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT are new operators defined as

Njk{Njnk,if jkNknj,if j<k.subscript𝑁𝑗𝑘casessubscript𝑁𝑗subscript𝑛𝑘if jk𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒subscript𝑁𝑘subscript𝑛𝑗if j<k𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒\displaystyle N_{jk}\equiv\begin{cases}N_{j}-n_{k},\quad\text{if $j\geq k$}\\ N_{k}-n_{j},\quad\text{if $j<k$}.\end{cases}italic_N start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT ≡ { start_ROW start_CELL italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_n start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , if italic_j ≥ italic_k end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_N start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , if italic_j < italic_k . end_CELL start_CELL end_CELL end_ROW (20)

From Eqs. 3.2, we see that the commutation relations of the traid anyon operators reduce to the bosonic case when acting on the same lattice site. However for operators acting on different sites, the commutation relations are either fermionic or bosonic, depending on the configuration of all the other particles on the lattice.

The traid anyon lattice model (14) generalizes to arbitrary abelian representations of the traid group in the following way:

H𝐻\displaystyle Hitalic_H =\displaystyle== JjL(bj+1eiπF(Nj)nj+1bj+h.c.)+U2jLnj(nj1),\displaystyle-J\sum_{j}^{L}\left(b^{\dagger}_{j+1}e^{i\pi F(N_{j})n_{j+1}}b_{j% }+h.c.\right)+\frac{U}{2}\sum^{L}_{j}n_{j}(n_{j}-1),- italic_J ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT ( italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_π italic_F ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) italic_n start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_h . italic_c . ) + divide start_ARG italic_U end_ARG start_ARG 2 end_ARG ∑ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - 1 ) , (21)

where F(Nj)𝐹subscript𝑁𝑗F(N_{j})italic_F ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) is a polynomial of Njsubscript𝑁𝑗N_{j}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT (15) that takes on even or odd integer values for integer inputs. F(Nj)𝐹subscript𝑁𝑗F(N_{j})italic_F ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) differs for each abelian representation and takes on the simplest form for the alternating cases: Falt(Nj)=Nj+Isubscript𝐹altsubscript𝑁𝑗subscript𝑁𝑗𝐼F_{\text{alt}}(N_{j})=N_{j}+Iitalic_F start_POSTSUBSCRIPT alt end_POSTSUBSCRIPT ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_I. For example, if one wanted to construct a model to realize the representation (++)(++-)( + + - ) for 4 traid anyons, we would need to find a polynomial that takes on the values F++(0)=0subscript𝐹absent00F_{++-}(0)=0italic_F start_POSTSUBSCRIPT + + - end_POSTSUBSCRIPT ( 0 ) = 0, F++(1)=0subscript𝐹absent10F_{++-}(1)=0italic_F start_POSTSUBSCRIPT + + - end_POSTSUBSCRIPT ( 1 ) = 0 and F++(2)=1subscript𝐹absent21F_{++-}(2)=1italic_F start_POSTSUBSCRIPT + + - end_POSTSUBSCRIPT ( 2 ) = 1 (or equivalently other even/odd integer values). In practice, for our numerical simulations in Sec. 4, we do not construct the polynomial explicitly and only set the desired inputs/outputs.

Analogously, we can find a Jordan-Wigner-like transformation that brings the generalized Hamiltonian of the traid anyon model (21) to the form (11) with a quadratic nearest neighbor tunneling term:

tjsubscript𝑡𝑗\displaystyle t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT \displaystyle\equiv eiπf(Nj)bj,superscript𝑒𝑖𝜋𝑓subscript𝑁𝑗subscript𝑏𝑗\displaystyle e^{-i\pi f(N_{j})}b_{j},italic_e start_POSTSUPERSCRIPT - italic_i italic_π italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (22)

where f(Nj)𝑓subscript𝑁𝑗f(N_{j})italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) is a polynomial of Njsubscript𝑁𝑗N_{j}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT (15), taking on odd or even integer values depending on the traid group representation which is to be realized. In particular, f(Nj+1)f(Nj)=F(Nj)nj+1𝑓subscript𝑁𝑗1𝑓subscript𝑁𝑗𝐹subscript𝑁𝑗subscript𝑛𝑗1f(N_{j+1})-f(N_{j})=F(N_{j})n_{j+1}italic_f ( italic_N start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT ) - italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = italic_F ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) italic_n start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT should be fulfilled, so that the kinetic term in Eq. 21 is quadratic when expressed in terms of these general traid anyon operators. For example, to construct traid anyon operators for the representation (++)(++-)( + + - ), we can set f(0)=f(1)=f(2)=0𝑓0𝑓1𝑓20f(0)=f(1)=f(2)=0italic_f ( 0 ) = italic_f ( 1 ) = italic_f ( 2 ) = 0 and f(3)=1𝑓31f(3)=1italic_f ( 3 ) = 1. It is easy to check that this yields the desired pattern of tunneling phases, provided we exclude three-body coincidences. Similarly, the generalized transformation (22) leads to generalized commutations relations for the tjsubscript𝑡𝑗t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT.

We note that the transformation defined in Eqs. 3.2 for the alternating abelian representations does not match the general form given by Eq. 22: the operator Mjsubscript𝑀𝑗M_{j}italic_M start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is not just a polynomial of Njsubscript𝑁𝑗N_{j}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT but also depends on the configuration of the particles to the left of site j𝑗jitalic_j (and not just their number). For instance, a two-body coincidence to the left of site j𝑗jitalic_j would result in an extra factor of 11-1- 1 in eiπMjsuperscript𝑒𝑖𝜋subscript𝑀𝑗e^{-i\pi M_{j}}italic_e start_POSTSUPERSCRIPT - italic_i italic_π italic_M start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. These potential additional factors of 11-1- 1 cancel out for terms of the form tj+1tjsuperscriptsubscript𝑡𝑗1subscript𝑡𝑗t_{j+1}^{\dagger}t_{j}italic_t start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT or nj=tjtjsubscript𝑛𝑗superscriptsubscript𝑡𝑗subscript𝑡𝑗n_{j}=t_{j}^{\dagger}t_{j}italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and have thus no influence on the Hamiltonian in Eq. 16. On the other hand, for observables like the two-point correlation function titjexpectationsuperscriptsubscript𝑡𝑖subscript𝑡𝑗\braket{t_{i}^{\dagger}t_{j}}⟨ start_ARG italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩, it can indeed make a difference, e.g. if there is a two-body coincidence between lattice sites i𝑖iitalic_i and j𝑗jitalic_j. The factor of 11-1- 1 then only occurs for one of the traid anyon operators and does not cancel out. The advantage of using the traid anyon operators defined in Eqs. 3.2 for the alternating abelian representations instead of the general form given by Eq. 22, is that the transformation to boson operators always results in the conceptually simple Hamiltonian in Eq. 14, even if we do not enforce a three-body hard-core constraint.

Finally, we would like to point out that the representation of traid anyons using bosons with number dependent tunneling matrix elements resembles the composite-particle picture of braid anyons in 2D (see, e.g. Ref. [6]). In the latter case, 2D braid anyons are mapped to charged bosons (or fermions) to which a magnetic flux is attached to implement the geometric phases picked up when particles are moved around each other. In the former case of our traid anyon lattice model, the number dependent tunneling matrix elements also give rise to generalized magnetic fluxes in the discrete configuration space of our system that are determined by the position of the particles and implement the geometric phases associated with the exchange of traid anyons; [See e.g. Figs. 5(b) and 6(e)].

4 Ground-state properties and signatures of fractional exclusion statistics

After introducing our lattice model for traid anyons above, we investigate its ground state properties in this section. To that end we numerically calculate the particle densities, energies and natural orbitals via exact diagonalization of Eqs. 14 and 21 (supplemented by DMRG simulations for higher filling factors in Fig. 9). These results show how the abelian traid group representations lie between fermions and bosons, and also hint at an intriguing connection to Haldane’s fractional exclusion statistics [23] (c.f. App. D).

For all our simulations in this section, we set the on-site interactions to zero (U=0𝑈0U=0italic_U = 0). We also enforced a three-body hard-core constraint to reduce computation times and to always preserve the characteristic pattern of exchange phases, which breaks down for tunneling processes involving three or more particles in our traid anyon lattice model (14). Because we stay in the low-energy, dilute regime for most of our simulations, three-body processes can be neglected for the calculated observables (see also the discussion in section 3.2). Consequently, the signatures of traid anyons reported below do not rely on this additional constraint.

4.1 Particle density: generalized Friedel oscillations

Refer to caption
Figure 7: Ground state particle densities for the traid anyon lattice model (14) for 3-6 particles on 20 lattice sites and U=0𝑈0U=0italic_U = 0. Each plot shows the densities corresponding to the two possible abelian representations of the traid group with alternating exchange phases of a given number of particles, i.e. the blue lines correspond to the I=0𝐼0I=0italic_I = 0 model where the first and second particles exchange like bosons, the second and third like fermions and so on. The exchange phases for I=1𝐼1I=1italic_I = 1 are flipped.
Refer to caption
Figure 8: Ground state particle densities for the generalized traid anyon lattice model (21) on 20 lattice sites for U=0𝑈0U=0italic_U = 0. Shown are all 8 possible abelian representations of the traid group for a system of 4 particles.

As a first physical observable, we consider the ground state particle densities. The particle density is independent of whether we formulate the model in terms of boson or traid anyon operators: nj=bjbj=tjtjexpectationsubscript𝑛𝑗expectationsuperscriptsubscript𝑏𝑗subscript𝑏𝑗expectationsuperscriptsubscript𝑡𝑗subscript𝑡𝑗\braket{n_{j}}=\braket{b_{j}^{\dagger}b_{j}}=\braket{t_{j}^{\dagger}t_{j}}⟨ start_ARG italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ = ⟨ start_ARG italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ = ⟨ start_ARG italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩, i.e. the number of bosons and traid anyons on each lattice site is identical. Fig. 7 shows the ground state densities of 3 to 6 traid anyons on 20202020 lattice sites. We depict the two possible abelian representations with alternating exchange phases (+++-+-...+ - + - …) and (++-+-+...- + - + …) for each number of particles. As expected, the reflection symmetry of the densities about the midpoint of the lattice embodies the symmetry of the arrangement of the exchange phases: While the individual density profiles for the representations with I=0𝐼0I=0italic_I = 0 and I=1𝐼1I=1italic_I = 1 are not symmetric for odd N𝑁Nitalic_N, they are mirror images of each other. In turn for even particle numbers, we find distinct, individually symmetric densities.

Examples of non-alternating abelian representations of the traid group are shown in Fig. 8, where the densities of all 8888 possible abelian representations (not just the alternating case) for 4444 particles are represented. The top left plot depicts the trivial case of bosons and pseudo-fermions, where all exchange phases are identical. The bottom two plots show 4 representations which can only be realized by the generalized version of our lattice model (21).

The density profiles depicted in Figs. 7 and 8 show that for each minus sign in the chosen traid-group representation, there is a minimum separating two local density maxima. Intuitively, depending on the exchange phases on both sides of the minus sign, these maxima are associated with either single particles resembling fermions, or two particles forming a ‘bosonic’ pair (separated by fermionic exchanges from its neighbors). For non-alternating traid group representations, when there are several plus signs next to each other, also ‘bosonic’ triples, quadruples, etc. can be observed (see Fig. 8). In the case of pseudo-fermions, i.e. the representations ()(---\cdots)( - - - ⋯ ) with only minus signs (see the top left plot in Fig. 8), these oscillations of the density correspond to Friedel oscillations [29], as they have been predicted also for the braid-anyon-Hubbard model [13]. These oscillations are a result of the fermionic two-particle correlations (reflecting Pauli exclusion), as they become visible in the density distribution close to a local defect, which is here given by the edge of our finite system with open boundary conditions. Thus, for the non-trivial representations, containing both plus and minus signs, the observed oscillations can be viewed as generalized Friedel oscillations.

4.2 Ground state energy and chemical potential

Refer to caption
Figure 9: Change in ground state energy (chemical potential) in units of J𝐽Jitalic_J after adding the N𝑁Nitalic_Nth particle to a system with N1𝑁1N-1italic_N - 1 particles on 30303030 lattice sites and no on-site interaction (U=0𝑈0U=0italic_U = 0). Traid anyons (14) with alternating exchange phases (+++-+-...+ - + - …) or (++-+-+...- + - + …) are compared to fermions, pseudo-fermions, bosons, and the braid-anyon-Hubbard model with θ=π2𝜃𝜋2\theta=\frac{\pi}{2}italic_θ = divide start_ARG italic_π end_ARG start_ARG 2 end_ARG. Except for bosons and fermions, a maximum site occupation of 2 particles was enforced.

Like the particle densities discussed above, the ground-state energy (as well as the full spectrum) does not depend on whether we consider bosons with number-dependent tunneling phases or traid anyons, since the Hamiltonians for both pictures are equivalent due to the construction of the transformation between them (3.2). We numerically calculate the ground state energies E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT of our lattice model and consider the zero-temperature chemical potential, i.e. the ground state energy difference E0(N)E0(N1)subscript𝐸0𝑁subscript𝐸0𝑁1E_{0}(N)-E_{0}(N-1)italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_N ) - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_N - 1 ) of a system with N𝑁Nitalic_N and N1𝑁1N-1italic_N - 1 respective particles. Note that this quantity is well defined only for a specific rule for how the sign τNsubscript𝜏𝑁\tau_{N}italic_τ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT is chosen, when the (N+1)𝑁1(N+1)( italic_N + 1 )st particle is added to the system.

The chemical potential directly provides a perspective on the relation between exchange statistics and exclusion statistics (see Appendix D below). Free bosons and free fermions (where by “free” we mean non-interacting) are the extreme cases. Any number of bosons can occupy the ground state, so each additional particle requires the same amount of energy, whereas each additional fermion must occupy the next higher single-particle energy eigenstate. Free particles with fractional exclusions statistics, where single-particle states can be occupied by a well defined finite number of particles, would, in a similar fashion, give rise to a clear fingerprint in the behaviour of the chemical potential as a function of the total particle number. Below, we will, indeed, see traces of such behaviour, but also deviations from it, which will be attributed to the fact that, as a result of the number-dependent tunneling phases, even for U=0𝑈0U=0italic_U = 0 the traid-anyon Hubbard model does not describe free particles.

We show the results for a system with 30303030 lattice sites, up to 10101010 particles and no on-site interactions (U=0𝑈0U=0italic_U = 0) in Fig. 9. We compare the two realizations of traid anyons with alternating exchange phases (14) to (spinless) fermions, pseudo-fermions, bosons, and to the braid-anyon-Hubbard model (13) with exchange phase θ=π2𝜃𝜋2\theta=\frac{\pi}{2}italic_θ = divide start_ARG italic_π end_ARG start_ARG 2 end_ARG.

The ground state energies of free fermions and free bosons can be calculated analytically from the single-particle energies E(k)=2Jcos(dk)𝐸𝑘2𝐽𝑑𝑘E(k)=-2J\cos(dk)italic_E ( italic_k ) = - 2 italic_J roman_cos ( italic_d italic_k ) of a system with M𝑀Mitalic_M lattice sites, lattice spacing d𝑑ditalic_d, and allowed quasi-momenta k=πνd(M+1)𝑘𝜋𝜈𝑑𝑀1k=\frac{\pi\nu}{d(M+1)}italic_k = divide start_ARG italic_π italic_ν end_ARG start_ARG italic_d ( italic_M + 1 ) end_ARG with ν=1,2,,M𝜈12𝑀\nu=1,2,...,Mitalic_ν = 1 , 2 , … , italic_M. For the bosonic ground state all bosons occupy the single-particle ground-state, thus E0(N)E0(N1)=subscript𝐸0𝑁subscript𝐸0𝑁1absentE_{0}(N)-E_{0}(N-1)=italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_N ) - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_N - 1 ) = 2Jcos(π/(M+1))2𝐽𝜋𝑀1-2J\cos(\pi/(M+1))- 2 italic_J roman_cos ( italic_π / ( italic_M + 1 ) ). For (spinless) fermions on the other hand, the Pauli exclusion principle implies that E0(N)limit-fromsubscript𝐸0𝑁E_{0}(N)-italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_N ) - E0(N1)subscript𝐸0𝑁1E_{0}(N-1)italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_N - 1 ) =2Jcos(Nπ/(M+1))absent2𝐽𝑁𝜋𝑀1=-2J\cos(N\pi/(M+1))= - 2 italic_J roman_cos ( italic_N italic_π / ( italic_M + 1 ) ). The ground state energies of the traid-anyon-Hubbard and braid-anyon-Hubbard model (13) were obtained numerically using exact diagonalization for N<7𝑁7N<7italic_N < 7 and DMRG for N7𝑁7N\geq 7italic_N ≥ 7. For completeness, we have also included pseudo-fermions corresponding to the ()(--\cdots-)( - - ⋯ - ) representation of the traid-anyon Hubbard model as well as to the braid-anyon-Hubbard model with θ=π𝜃𝜋\theta=\piitalic_θ = italic_π. Their chemical potential agrees nicely with that of free fermions in the dilute regime (roughly below quarter filling or N=7𝑁7N=7italic_N = 7), but as the density increases one can see a deviation because multiply-occupied sites are allowed for pseudo-fermions, unlike for true fermions.

We chose the braid-anyon-Hubbard model with exchange phase θ=π2𝜃𝜋2\theta=\frac{\pi}{2}italic_θ = divide start_ARG italic_π end_ARG start_ARG 2 end_ARG for comparison, because it provides a different notion of particles exhibiting behaviour ‘right in the middle’ between fermions and bosons: the anyons in the braid-anyon-Hubbard model always exchange with the same phase θ=π2𝜃𝜋2\theta=\frac{\pi}{2}italic_θ = divide start_ARG italic_π end_ARG start_ARG 2 end_ARG, averaging between bosonic and fermionic behavior. In contrast, traid anyons in the representations (++)(+-+-...)( + - + - … ) or (++)(-+-+...)( - + - + … ) alternate between exchanging fermion-like (corresponding to an exchange phase θ=π𝜃𝜋\theta=\piitalic_θ = italic_π) and boson-like (θ=0𝜃0\theta=0italic_θ = 0). These different notions of interpolation between fermions and bosons are also reflected in the chemical potentials in Fig. 9. While the chemical potential lies approximately in the middle of the fermionic and bosonic case for both traid anyons and the braid-anyon-Hubbard model with θ=π2𝜃𝜋2\theta=\frac{\pi}{2}italic_θ = divide start_ARG italic_π end_ARG start_ARG 2 end_ARG, the behaviour is nonetheless strikingly different: We observe a smooth increase of the potential for the braid-anyon-Hubbard model, while the traid anyons exhibit a step-like behaviour, where the change in the chemical potential increases or decreases in an alternating fashion.

The step-wise increase of the chemical potential, observed for the staggered traid anyons, can be interpreted as a signature of approximate Haldane-like semionic fractional exclusion statistics, i.e. of a generalized Pauli exclusion principle, where single-particle states can be occupied by up to two particles (c.f. App. D). Ideally, such behaviour might be associated with the chemical potential increasing for every other particle added and staying constant otherwise. We, however, find that rather than staying constant in the latter case, the chemical potential decreases. This hints at an effectively induced attractive interaction, something previously noted for the braid-anyon-Hubbard model [30]. In the next subsection, where we investigate the natural orbitals and their occupations, we find further evidence for the emergence of approximate fractional exclusion statistics and an explanation for the observed decrease of the chemical potential in every other step.

The emergence of semionic exclusion statistics can be understood intuitively as a consequence of the staggered exchange statistics of the traid anyons. Namely, when the rightmost particles form a ‘bosonic’ pair, a newly added particle comes with a minus sign and thus plays the role of a new single ‘fermion’ on the right edge. In turn, when the rightmost particle resembles a single ‘fermion’, the added particle comes with a plus sign, so that it can occupy the same space as the former rightmost particle (both form a new ‘bosonic’ pair).

It is straightforward to generalize these ideas to non-staggered representations. We then expect to find scenarios where 3 (or more) neighboring particles behave like bosons to each other, e.g. for (++++)(++-++-...)( + + - + + - … ). Also mixed fractional statistics can be imagined, e.g. for (++++)(+-++-+-...)( + - + + - + - … ) pairs and triples of particles are expected to be in the same state. Such cases will be discussed in the next subsection.

4.3 Natural orbitals

Refer to caption
Figure 10: Ground state natural orbitals (insets) and their occupation numbers (main plots) for our traid anyon lattice model (14, 21) on 20202020 lattice sites without on-site interaction (U=0𝑈0U=0italic_U = 0), obtained from diagonalizing the single-particle density matrix titjexpectationsuperscriptsubscript𝑡𝑖subscript𝑡𝑗\braket{t_{i}^{\dagger}t_{j}}⟨ start_ARG italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ with the traid anyon operators defined in Eqs. 3.2 and 22.

The natural orbitals are defined as the eigenvectors of the single-particle density matrix (SPDM) cicjdelimited-⟨⟩superscriptsubscript𝑐𝑖subscript𝑐𝑗\langle c_{i}^{\dagger}c_{j}\rangle⟨ italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⟩, where cisubscriptsuperscript𝑐𝑖c^{\dagger}_{i}italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and cjsubscript𝑐𝑗c_{j}italic_c start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT are generic creation and annihilation operators, respectively. The corresponding eigenvalues are the natural orbital’s occupation numbers that represent how many particles occupy each natural orbital on average. For free the many-body ground state of free bosons (ci=bisubscript𝑐𝑖subscript𝑏𝑖c_{i}=b_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT) the single eigenvector with non-zero eigenvalue of the SPDM would be the state-space vector of the single-particle ground state and the corresponding eigenvalue would give the occupation N𝑁Nitalic_N of that state. For free fermions, one finds N𝑁Nitalic_N eigenvectors with non-zero eigenvalues of 1111, corresponding to the singly-occupied single-particle energy eigenstates forming the Fermi sea. Considering interacting bosonic or fermionic systems, the natural orbitals no longer correspond to the single-particle energy eigenstates and the occupation numbers of the natural orbitals are generally not integer valued anymore. For free particles with fractional exclusion statistics, one would expect eigenvalues corresponding to occupations that take integer values larger than one, but smaller or equal than an allowed maximum occupation (e.g. two for so-called semions). Below, we indeed find approximately such behaviour, with small deviations that we attribute to the presence of interactions associated with the number-dependent tunneling even for U=0𝑈0U=0italic_U = 0.

In order to investigate the properties of the traid anyons, we need to consider the SPDM with respect to the traid-anyon operators,

titj=bieiπ(MiMj)bjexpectationsuperscriptsubscript𝑡𝑖subscript𝑡𝑗expectationsuperscriptsubscript𝑏𝑖superscript𝑒𝑖𝜋subscript𝑀𝑖subscript𝑀𝑗subscript𝑏𝑗\displaystyle\braket{t_{i}^{\dagger}t_{j}}=\braket{b_{i}^{\dagger}e^{i\pi\left% (M_{i}-M_{j}\right)}b_{j}}⟨ start_ARG italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ = ⟨ start_ARG italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_π ( italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_M start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ (23)

Note that it is different from the bosonic one, titjbibjexpectationsuperscriptsubscript𝑡𝑖subscript𝑡𝑗expectationsuperscriptsubscript𝑏𝑖subscript𝑏𝑗\braket{t_{i}^{\dagger}t_{j}}\neq\braket{b_{i}^{\dagger}b_{j}}⟨ start_ARG italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ ≠ ⟨ start_ARG italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩. An intuitive understanding of how the occupation numbers for hypothetical free traid anyons would look like a given traid group representation (τ1,,τN1)subscript𝜏1subscript𝜏𝑁1(\tau_{1},\ldots,\tau_{N}-1)( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_τ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT - 1 ) can be gained as follows: Whenever there is a string of τi=+1subscript𝜏𝑖1\tau_{i}=+1italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = + 1 of length m𝑚mitalic_m in a representation, then those m+1𝑚1m+1italic_m + 1 particles exchange like bosons and can all occupy the same state. Whenever there is τi=1subscript𝜏𝑖1\tau_{i}=-1italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = - 1 in a representation, a new orbital must be added to account for the antisymmetric exchange. For example with N=3𝑁3N=3italic_N = 3, the alternating representation (+)(+-)( + - ) would have one orbital localized on the left with occupation number 2222 corresponding to the two particles with boson-like exchanges, and one orbital with occupation number 1111 localized on the right side. The representation (+)(-+)( - + ) would have the same occupation numbers as (+)(+-)( + - ) but mirrored orbitals. For N=4𝑁4N=4italic_N = 4, the ground state of representation (+)(-+-)( - + - ) would have three orbitals, one with occupation number 2222 localized in the center, and two with occupation number 1111 localized on the sides, whereas (++)(+-+)( + - + ) would have two orbitals with occupation number 2222 localized on either side. Generally, all alternating representations will have a maximum occupancy of 2222, i.e., semionic exclusion statistics. The reasoning presented in this paragraph has to be taken with a grain of salt, as the actual natural orbitals are delocalized (as is know for instance for the case of free fermions). But it provides an intuitive understanding of the expected fractional exclusion statistics, as we indeed find it in our numerical results below (up to corrections which we attribute to the presence of interactions even for U=0𝑈0U=0italic_U = 0 associated with the number dependent tunneling).

Our numerical results are depicted in Fig. 10. We display the occupation numbers of the natural orbitals (main plots), as well as the one-particle densities of the first three natural orbitals ordered by their occupation numbers (insets). While Figs. 10(a-d) show examples of traid anyons with alternating hopping signs, specified by Eqs. 14 and 17a, Figs. 10(e,f) additionally show two different nontrivial abelian representations for 4 particles, namely (+)(+--)( + - - ) and (++)(++-)( + + - ), defined by the generalized traid anyon lattice model (21, 22). In analogy to the ground state densities and energies discussed above, the densities of the natural orbitals corresponding to mirrored abelian representations [i.e., (+)(+-)( + - ) and (+)(-+)( - + )] are also mirrored with identical occupation numbers.

The most striking feature of the natural orbitals in the traid anyon picture is their (near) integer-valued occupation numbers that approximately conform to the free traid-anyon hypothesis described above. In particular, for the three-particle case shown in Fig. 10(a), we find occupation numbers of 1.984 and 1.009 for the first two natural orbitals. As we increase N𝑁Nitalic_N, in Fig 10(b)-(d), the expected pattern remains, although there are larger deviations from the predicted integer values. This can be interpreted as an (approximate) constraint of a maximum of two traid anyons occupying each natural orbital for the alternating representations. This behaviour provides further indication of an approximate emergent Haldane-like fractional semionic exclusion statistics in the traid-anyon-Hubbard model (14), as we discuss in Sec. 4.2 and App. D. We note that other (i.e. non-staggered) traid representations can also lead to a different generalized exclusion statistics, as exemplified in Fig. 10(e) and (f), where the occupation numbers again approximately conform to the integer values of the free traid anyon hypothesis.

The squared absolute value of the natural orbitals, as they are shown in the insets of Fig. 10, provide further insight into how the ground-state density distributions shown in Figs. 7 and 8 are formed. The shape of the natural orbitals also provides a possible explanation for the observation that the chemical potential in Fig. 9 decreases every other time when a particle is added to the system to form a ‘bosonic’ pair with a previously fermion-like single particle. Namely, in Fig. 10, one sees that the shape of the natural orbitals depends on the number of particles in the system. In particular, we find that the width of the natural orbitals tends to increase with their occupation, so that the kinetic energy is lowered for multiply-occupied orbitals relative to that of singly-occupied ones, suggesting a mechanism for the reduction of the chemical potential. This behaviour is clearly different from that found for the ground state of both free bosons and fermions, for which the natural orbitals are not number dependent and simply given by the single-particle eigenstates.

If this broadening of multiply-occupied orbitals can be interpreted as a signature of some form of (effective) attractive interactions in the system, it might also explain the deviations from perfectly quantized occupations of the natural orbitals visible in Fig. 10. The origin of these interactions must be related to the form of the commutation relations of the traid operators (3.2). Different from the standard bosonic and fermionic ones, they involve ‘interaction’ terms, i.e. terms that involve products of more than two annihilation and creation operators. One immediate consequence is, that the commutation relations become basis-dependent. That is, transformations tα=iα|itisubscript𝑡𝛼subscript𝑖inner-product𝛼𝑖subscript𝑡𝑖t_{\alpha}=\sum_{i}\braket{\alpha}{i}t_{i}italic_t start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟨ start_ARG italic_α end_ARG | start_ARG italic_i end_ARG ⟩ italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT to a basis of single particle states |αket𝛼|\alpha\rangle| italic_α ⟩ (for instance the basis of natural orbitals), other than the local Wannier basis |iket𝑖|i\rangle| italic_i ⟩ defining the lattice sites, change the form of the commutation relations.

5 Continuum limit

In section 3 we constructed a lattice model (14) using bosons with non-local, density-dependent Peierls phases that corresponds to abelian representations of the traid group. Here we derive the continuum limit of this model with staggered exchange phases (14). This approach can be easily generalized also to other abelian representations of the traid group. Similar to the Tonks-Girardeau model in which one-dimensional fermions can be mapped to hard-core bosons, we find that in the continuum limit abelian traid anyons can be expressed in terms of bosons with hard-core contact interactions. Importantly, these interactions depend on the ordinality of the particles and are, therefore, non-local.

We find that the eigenstates of the continuum model can be directly mapped to traid-anyon wave-functions, as they were constructed previously from general considerations [16, 17]. This mapping is similar to that from hard-core bosons to fermions. Namely, in our model, hard-core interactions occur whenever two particles meet and exchange like fermions. As a result, the many-body wave-function becomes zero at this coincidence, allowing to flip the sign of the wave function, when mapping between bosons and traid anyons, without increasing the energy. The fact that we recover previous results from the continuum limit of our model system provides an a-posteriori justification of the construction of our model.

5.1 Continuum limit of the lattice model

We take the continuum limit of our lattice model by letting the lattice spacing go to zero,

d0,𝑑0d\to 0,italic_d → 0 , (24)

while keeping the physical length l=Ld𝑙𝐿𝑑l=Lditalic_l = italic_L italic_d constant. We will also see that the effective mass of the continuum particles m*superscript𝑚m^{*}italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT is defined by Jd2=2/(2m*)𝐽superscript𝑑2superscriptPlanck-constant-over-2-pi22superscript𝑚Jd^{2}=\hbar^{2}/(2m^{*})italic_J italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 2 italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) 333This result follows also directly from expanding the dispersion relation of free bosons with respect to the lattice spacing, ε(k)=2Jcos(dk)2J+Jd2k22k22m*+const.𝜀𝑘2𝐽𝑑𝑘similar-to-or-equals2𝐽𝐽superscript𝑑2superscript𝑘2superscriptPlanck-constant-over-2-pi2superscript𝑘22superscript𝑚const.\varepsilon(k)=-2J\cos(dk)\simeq-2J+Jd^{2}k^{2}\equiv\frac{\hbar^{2}k^{2}}{2m^% {*}}+\text{const.}italic_ε ( italic_k ) = - 2 italic_J roman_cos ( italic_d italic_k ) ≃ - 2 italic_J + italic_J italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≡ divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG + const.. and that the Hubbard interactions of strength U𝑈Uitalic_U give rise to a contact interaction potential gδ(x)𝑔𝛿𝑥g\delta(x)italic_g italic_δ ( italic_x ) of strength g=Ud𝑔𝑈𝑑g=Uditalic_g = italic_U italic_d. Thus, we have to scale L𝐿Litalic_L, J𝐽Jitalic_J, and U𝑈Uitalic_U like

L=ld,J=22m*1d2,U=gd,formulae-sequence𝐿𝑙𝑑formulae-sequence𝐽superscriptPlanck-constant-over-2-pi22superscript𝑚1superscript𝑑2𝑈𝑔𝑑L=\frac{l}{d},\qquad J=\frac{\hbar^{2}}{2m^{*}}\frac{1}{d^{2}},\qquad U=\frac{% g}{d},italic_L = divide start_ARG italic_l end_ARG start_ARG italic_d end_ARG , italic_J = divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , italic_U = divide start_ARG italic_g end_ARG start_ARG italic_d end_ARG , (25)

with m*superscript𝑚m^{*}italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT and g𝑔gitalic_g held constant. Moreover, we have to replace sums by integrals,

j=1L1d0l𝑑x,superscriptsubscript𝑗1𝐿1𝑑superscriptsubscript0𝑙differential-d𝑥\sum_{j=1}^{L}\to\frac{1}{d}\int_{0}^{l}\!dx,∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT → divide start_ARG 1 end_ARG start_ARG italic_d end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT italic_d italic_x , (26)

and bosonic annihilation and creation operators bjsubscript𝑏𝑗b_{j}italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT by bosonic field operators ψ(x)𝜓𝑥\psi(x)italic_ψ ( italic_x ), according to [31]

ψ()(x)bj()d,superscript𝜓𝑥subscriptsuperscript𝑏𝑗𝑑\displaystyle\psi^{(\dagger)}(x)\to\frac{b^{(\dagger)}_{j}}{\sqrt{d}},italic_ψ start_POSTSUPERSCRIPT ( † ) end_POSTSUPERSCRIPT ( italic_x ) → divide start_ARG italic_b start_POSTSUPERSCRIPT ( † ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG italic_d end_ARG end_ARG , (27)

where [ψ(x),ψ(x)]=δ(xx)𝜓𝑥superscript𝜓superscript𝑥𝛿𝑥superscript𝑥[\psi(x),\psi^{\dagger}(x^{\prime})]=\delta(x-x^{\prime})[ italic_ψ ( italic_x ) , italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] = italic_δ ( italic_x - italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ).

It is straightforward to obtain the continuum limit of the on-site Hubbard interactions

Hos=U2jbjbjbjbjsubscript𝐻os𝑈2subscript𝑗superscriptsubscript𝑏𝑗superscriptsubscript𝑏𝑗subscript𝑏𝑗subscript𝑏𝑗\displaystyle H_{\text{os}}=\frac{U}{2}\sum_{j}b_{j}^{\dagger}b_{j}^{\dagger}b% _{j}b_{j}italic_H start_POSTSUBSCRIPT os end_POSTSUBSCRIPT = divide start_ARG italic_U end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT \displaystyle\to g2𝑑xψ(x)ψ(x)ψ(x)ψ(x)=g2𝑑x:ρ2(x):,:𝑔2differential-d𝑥superscript𝜓𝑥superscript𝜓𝑥𝜓𝑥𝜓𝑥𝑔2differential-d𝑥superscript𝜌2𝑥:absent\displaystyle\frac{g}{2}\int\!dx\,\psi^{\dagger}(x)\psi^{\dagger}(x)\psi(x)% \psi(x)=\frac{g}{2}\int\!dx\,:\rho^{2}(x):,divide start_ARG italic_g end_ARG start_ARG 2 end_ARG ∫ italic_d italic_x italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_x ) italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_x ) italic_ψ ( italic_x ) italic_ψ ( italic_x ) = divide start_ARG italic_g end_ARG start_ARG 2 end_ARG ∫ italic_d italic_x : italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x ) : , (28)

where we have introduced the density operator ρ(x)=ψ(x)ψ(x)𝜌𝑥superscript𝜓𝑥𝜓𝑥\rho(x)=\psi^{\dagger}(x)\psi(x)italic_ρ ( italic_x ) = italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_x ) italic_ψ ( italic_x ) as well as the usual convention that :::::\bullet:: ∙ : means normal ordering of all operators between the colons, i.e. all creation operators are positioned to the left of the annihilation operators.

Because of the number-dependent Peierls phases, the continuum limit of the tunneling term in Eq. (14) is more involved. Restricting ourselves to the case I=0𝐼0I=0italic_I = 0, it reads

Htun=JjL(bj+1eiπNjnj+1bj+h.c.).H_{\text{tun}}=-J\sum_{j}^{L}\left(b^{\dagger}_{j+1}e^{i\pi N_{j}n_{j+1}}b_{j}% +\mathrm{h.c.}\right).italic_H start_POSTSUBSCRIPT tun end_POSTSUBSCRIPT = - italic_J ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT ( italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_π italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + roman_h . roman_c . ) . (29)

At the end, we will reintroduce the dependence on I𝐼Iitalic_I, simply via the substitution NjNj+Isubscript𝑁𝑗subscript𝑁𝑗𝐼N_{j}\to N_{j}+Iitalic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT → italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_I. Calculating the continuum limit of the braid-anyon-Hubbard model, Bonkhoff et al. managed to preserve the crucial symmetry of the anyonic exchange phase under the transformation θθ+2π𝜃𝜃2𝜋\theta\to\theta+2\piitalic_θ → italic_θ + 2 italic_π in the continuum, by normal ordering the operators in the complex phase factor before taking the limit [15]. We closely follow their approach in deriving the continuum limit of the traid anyon lattice model. In our case, we want to preserve the symmetry of the complex phase factor under the transformation NiNi+2subscript𝑁𝑖subscript𝑁𝑖2N_{i}\to N_{i}+2italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT → italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + 2. Thus, we calculate the continuum limit of the ‘number-to-the-left’ operator Nisubscript𝑁𝑖N_{i}italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT independently and then write all other operators in the number dependent hopping phase in normal order before taking the limit. The continuum limit of this operator is given by

Nj=kjbkbksubscript𝑁𝑗subscript𝑘𝑗subscriptsuperscript𝑏𝑘subscript𝑏𝑘\displaystyle N_{j}=\sum_{k\leq j}b^{\dagger}_{k}b_{k}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_k ≤ italic_j end_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT \displaystyle\to 0x𝑑xψ(x)ψ(x)=0xdxρ(x)N(x),superscriptsubscript0𝑥differential-dsuperscript𝑥superscript𝜓superscript𝑥𝜓superscript𝑥superscriptsubscript0𝑥differential-dsuperscript𝑥𝜌superscript𝑥𝑁𝑥\displaystyle\int_{0}^{x}\!dx^{\prime}\,\psi^{\dagger}(x^{\prime})\psi(x^{% \prime})=\int_{0}^{x}\mathop{}\!\mathrm{d}x^{\prime}\rho(x^{\prime})\equiv N(x),∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT italic_d italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_ψ ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT roman_d italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_ρ ( italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ≡ italic_N ( italic_x ) , (30)

where xjd𝑥𝑗𝑑x\equiv jditalic_x ≡ italic_j italic_d. Analogous to the discrete case, the operator N(x)𝑁𝑥N(x)italic_N ( italic_x ) counts the number of particles to the left of position x𝑥xitalic_x. In particular, one has N(l)=N𝑁𝑙𝑁N(l)=Nitalic_N ( italic_l ) = italic_N.

Next, we expand the exponential phase factor:

Htun=Jjbj+1(q=0(iπ)qq!Njqnj+1q)bj+h.c..subscript𝐻tun𝐽subscript𝑗subscriptsuperscript𝑏𝑗1superscriptsubscript𝑞0superscript𝑖𝜋𝑞𝑞superscriptsubscript𝑁𝑗𝑞superscriptsubscript𝑛𝑗1𝑞subscript𝑏𝑗h.c.H_{\text{tun}}=-J\sum_{j}b^{\dagger}_{j+1}\left(\sum_{q=0}^{\infty}\frac{(i\pi% )^{q}}{q!}N_{j}^{q}n_{j+1}^{q}\right)b_{j}+\text{h.c.}.italic_H start_POSTSUBSCRIPT tun end_POSTSUBSCRIPT = - italic_J ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_q = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( italic_i italic_π ) start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT end_ARG start_ARG italic_q ! end_ARG italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT ) italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + h.c. . (31)

The powers of the number operators nj+1subscript𝑛𝑗1n_{j+1}italic_n start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT can now be written in the normal ordered form

nj+1q=m=0S(q,m)dm(bj+1d)m(bj+1d)m,superscriptsubscript𝑛𝑗1𝑞superscriptsubscript𝑚0𝑆𝑞𝑚superscript𝑑𝑚superscriptsubscriptsuperscript𝑏𝑗1𝑑𝑚superscriptsubscript𝑏𝑗1𝑑𝑚\displaystyle n_{j+1}^{q}=\sum_{m=0}^{\infty}S(q,m)d^{m}\left(\frac{b^{\dagger% }_{j+1}}{\sqrt{d}}\right)^{m}\left(\frac{b_{j+1}}{\sqrt{d}}\right)^{m},italic_n start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_m = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_S ( italic_q , italic_m ) italic_d start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( divide start_ARG italic_b start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG italic_d end_ARG end_ARG ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT ( divide start_ARG italic_b start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG italic_d end_ARG end_ARG ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT , (32)

with the Stirling numbers of the second kind S(q,m)𝑆𝑞𝑚S(q,m)italic_S ( italic_q , italic_m ) [32, 15]. We now take the continuum limit, approximate the field operators at x+d𝑥𝑑x+ditalic_x + italic_d as:

ψ(x+d)𝜓𝑥𝑑\displaystyle\psi(x+d)italic_ψ ( italic_x + italic_d ) similar-to-or-equals\displaystyle\simeq ψ(x)+dxψ(x)+d222x2ψ(x)ψ+dψx+d22ψxx,𝜓𝑥𝑑𝑥𝜓𝑥superscript𝑑22superscript2superscript𝑥2𝜓𝑥𝜓𝑑subscript𝜓𝑥superscript𝑑22subscript𝜓𝑥𝑥\displaystyle\psi(x)+d\frac{\partial}{\partial x}\psi(x)+\frac{d^{2}}{2}\frac{% \partial^{2}}{\partial x^{2}}\psi(x)\equiv\psi+d\psi_{x}+\frac{d^{2}}{2}\psi_{% xx},italic_ψ ( italic_x ) + italic_d divide start_ARG ∂ end_ARG start_ARG ∂ italic_x end_ARG italic_ψ ( italic_x ) + divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_ψ ( italic_x ) ≡ italic_ψ + italic_d italic_ψ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG italic_ψ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT , (33)

(note that higher order terms will not contribute in the continuum limit) and finally sort all resulting terms by their power in the lattice spacing d𝑑ditalic_d. The terms of zeroth order in d𝑑ditalic_d are

Jdx(ψψ+h.c.)=2JN.\displaystyle-J\int\mathop{}\!\mathrm{d}x\left(\psi^{\dagger}\psi+\mathrm{h.c.% }\right)=-2JN.- italic_J ∫ roman_d italic_x ( italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ψ + roman_h . roman_c . ) = - 2 italic_J italic_N . (34)

This amounts to a constant energy contribution, which we can neglect in the final continuum model.

For the first-order terms we find

2Jddx(cos(πN(x))1)ψψψψ=2m*dx1cos(πN(x))d:ρ2:.:2𝐽𝑑differential-d𝑥𝜋𝑁𝑥1superscript𝜓superscript𝜓𝜓𝜓superscriptPlanck-constant-over-2-pi2superscript𝑚differential-d𝑥1𝜋𝑁𝑥𝑑superscript𝜌2:absent\displaystyle-2Jd\int\mathop{}\!\mathrm{d}x(\cos(\pi N(x))-1)\psi^{\dagger}% \psi^{\dagger}\psi\psi=\frac{\hbar^{2}}{m^{*}}\int\mathop{}\!\mathrm{d}x\frac{% 1-\cos(\pi N(x))}{d}:\rho^{2}:.- 2 italic_J italic_d ∫ roman_d italic_x ( roman_cos ( italic_π italic_N ( italic_x ) ) - 1 ) italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ψ italic_ψ = divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG ∫ roman_d italic_x divide start_ARG 1 - roman_cos ( italic_π italic_N ( italic_x ) ) end_ARG start_ARG italic_d end_ARG : italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT : . (35)

This term describes a diverging (i.e. hard-core) two-body contact interaction for even values of N(x)𝑁𝑥N(x)italic_N ( italic_x ) due to the 1/d1𝑑1/d1 / italic_d dependence.

Finally, the second-order terms read

2m*dx[(1cos(πN(x))):ρ3:+12(ψxxψ+ψψxx)],-\frac{\hbar^{2}}{m^{*}}\int\!dx\Big{[}\left(1-\cos(\pi N(x))\right):\rho^{3}:% +\frac{1}{2}\left(\psi^{\dagger}_{xx}\psi+\psi^{\dagger}\psi_{xx}\right)\Big{]},- divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG ∫ italic_d italic_x [ ( 1 - roman_cos ( italic_π italic_N ( italic_x ) ) ) : italic_ρ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT : + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT italic_ψ + italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ) ] , (36)

corresponding to a kinetic energy term and a three-body interaction term. All other terms are of order 𝒪(d3)𝒪superscript𝑑3\mathcal{O}(d^{3})caligraphic_O ( italic_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) and vanish in the continuum limit.

Summing up all terms (and including the continuum limit of the on-site interaction term in Eq. 14a), we obtain the final continuum Hamiltonian

Hcont.=dx[22m*ψx2ψ+V2:ρ2:+V3:ρ3:],H_{\mathrm{cont.}}=\int\mathop{}\!\mathrm{d}x\left[-\frac{\hbar^{2}}{2m^{*}}% \psi^{\dagger}\partial_{x}^{2}\psi+V_{2}:\rho^{2}:+\enspace V_{3}:\rho^{3}:% \right],italic_H start_POSTSUBSCRIPT roman_cont . end_POSTSUBSCRIPT = ∫ roman_d italic_x [ - divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ + italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT : italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT : + italic_V start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT : italic_ρ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT : ] , (37)

with the two-body and three-body interaction coefficients

V2(x)subscript𝑉2𝑥\displaystyle V_{2}(x)italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x ) =\displaystyle== 2m*d(1cos[π(N(x)+I)])+g2,superscriptPlanck-constant-over-2-pi2superscript𝑚𝑑1𝜋𝑁𝑥𝐼𝑔2\displaystyle\frac{\hbar^{2}}{m^{*}d}\left(1-\cos\left[\pi(N(x)+I)\right]% \right)+\frac{g}{2},divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_d end_ARG ( 1 - roman_cos [ italic_π ( italic_N ( italic_x ) + italic_I ) ] ) + divide start_ARG italic_g end_ARG start_ARG 2 end_ARG , (38)
V3(x)subscript𝑉3𝑥\displaystyle V_{3}(x)italic_V start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_x ) =\displaystyle== 2m*(cos[π(N(x)+I)]1),superscriptPlanck-constant-over-2-pi2superscript𝑚𝜋𝑁𝑥𝐼1\displaystyle\frac{\hbar^{2}}{m^{*}}\big{(}\cos\left[\pi(N(x)+I)\right]-1\big{% )},divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG ( roman_cos [ italic_π ( italic_N ( italic_x ) + italic_I ) ] - 1 ) , (39)

where we reintroduced the index I𝐼Iitalic_I, determining which of the two abelian representations of the traid group with alternating exchange phases the model realizes.

From the two-body interaction coefficient V2subscript𝑉2V_{2}italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (38), we can see that two particles exchanging like fermions in the lattice model (i.e., where N(x)+I𝑁𝑥𝐼N(x)+Iitalic_N ( italic_x ) + italic_I is odd), now exhibit a hard-core interaction in the continuum limit. On the other hand, particles that exchange like bosons (N(x)+I𝑁𝑥𝐼N(x)+Iitalic_N ( italic_x ) + italic_I is even) do not interact at all with one another in the continuum model (if g=0𝑔0g=0italic_g = 0). The hard-core two-body interaction between alternating pairs of particles also implies a three-body hard-core constraint, which is characteristic of the traid group. Thus, it emerges naturally in the continuum limit of the traid anyon-Hubbard model (14), even if we do not impose it by additional potentials or constraints. As a result, in the d0𝑑0d\to 0italic_d → 0 limit, we can neglect the finite-strength three-body interaction term V3(x)subscript𝑉3𝑥V_{3}(x)italic_V start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_x ) in our alternating traid anyon model and simply write

Hcont.subscript𝐻cont\displaystyle H_{\mathrm{cont.}}italic_H start_POSTSUBSCRIPT roman_cont . end_POSTSUBSCRIPT =\displaystyle== dx[22m*ψx2ψ+V2(x):ρ2:].\displaystyle\int\mathop{}\!\mathrm{d}x\left[-\frac{\hbar^{2}}{2m^{*}}\psi^{% \dagger}\partial_{x}^{2}\psi+V_{2}(x):\rho^{2}:\right].∫ roman_d italic_x [ - divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG italic_ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ + italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x ) : italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT : ] . (40)

The emergence of the contact interactions with infinite (zero) strength for two particles that behave like fermions (bosons) to each other is also consistent with previous results of Refs. [33, 12]. There it was shown for the braid-anyon-Hubbard model, that the combination of number dependent tunneling with anyonic phase θ𝜃\thetaitalic_θ and on-site Hubbard interactions U𝑈Uitalic_U gives rise to an effective contact interaction, whose strength diverges for θπ𝜃𝜋\theta\to\piitalic_θ → italic_π and vanishes when both θ=0𝜃0\theta=0italic_θ = 0 and U=0𝑈0U=0italic_U = 0.

Comparing the continuum limit (37) of the traid-anyon-Hubbard model to the continuum limit of the braid-anyon-Hubbard model [15], we see that we recover the same structure of the Hamiltonian, when we make the identification θπ(N(x)+I)𝜃𝜋𝑁𝑥𝐼\theta\to\pi(N(x)+I)italic_θ → italic_π ( italic_N ( italic_x ) + italic_I ). The crucial difference between the two models is that in the case of the traid anyons the coefficients V2subscript𝑉2V_{2}italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and V3subscript𝑉3V_{3}italic_V start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT are promoted to operators, resulting in non-local interactions between the particles. Moreover the current interaction term that arises in the continuum limit of the braid-anyon-Hubbard model vanishes for our model. Thus, in contrast to the continuum model of the braid-anyon-Hubbard model, the traid-anyon-Hubbard model constructed here gives rise to a Galilean invariant Hamiltonian.

A possible explanation for this difference is that the breaking of Galilean invariance via a dependence on the current operator originates from the requirement to distinguish exchange processes corresponding to the left particle moving over or under the right one, as depicted in Figs.2(b) and (c). While in the braid-anyon-Hubbard model these processes can be distinguished via different paths that form a non-contractable loop in the discrete configuration space [see Fig. 5], this is not possible in the continuum limit. Thus, the dependence on the current operator (also found in the chiral BF and Kundu models) can be interpreted as a way to “translate” the distinction between these two processes in the lattice where two particles exchange either via leftward or rightward tunneling to the continuum. A similar observation on the necessity for breaking Galilean invariance to implement fractional exchange statistics in 1D models has been made for excitons in the Haldane-Shastry model [34, 35]. However, for traid anyons this is not the case, suggesting that they correspond to a more natural (intrinsic) definition of anyons in the continuum in 1D, as is also suggested by topology of the continuum strand diagrams discussed above.

5.2 Continuum model in first quantization

From (37), we can derive the bosonic continuum model in first quantization. First consider the Hamiltonian restricted to the sector 𝒳12NNsubscript𝒳12𝑁superscript𝑁\mathcal{X}_{12\cdots N}\subset{\mathbb{R}}^{N}caligraphic_X start_POSTSUBSCRIPT 12 ⋯ italic_N end_POSTSUBSCRIPT ⊂ blackboard_R start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT where the particle labels coincide with particle order x1x2xNsubscript𝑥1subscript𝑥2subscript𝑥𝑁x_{1}\leq x_{2}\leq\cdots\leq x_{N}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ ⋯ ≤ italic_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT. The Hamiltonian that realizes the alternating representations characterized by I𝐼Iitalic_I has the form

Hcont.FQ=j(22m*2xj2+V~2(j)δ(xjxj+1)),superscriptsubscript𝐻contFQsubscript𝑗superscriptPlanck-constant-over-2-pi22superscript𝑚superscript2superscriptsubscript𝑥𝑗2subscript~𝑉2𝑗𝛿subscript𝑥𝑗subscript𝑥𝑗1H_{\mathrm{cont.}}^{\text{FQ}}=\sum_{j}\left(-\frac{\hbar^{2}}{2m^{*}}\frac{% \partial^{2}}{\partial x_{j}^{2}}+\tilde{V}_{2}(j)\delta(x_{j}-x_{j+1})\right),italic_H start_POSTSUBSCRIPT roman_cont . end_POSTSUBSCRIPT start_POSTSUPERSCRIPT FQ end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( - divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_j ) italic_δ ( italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT ) ) , (41a)
where the two-body interaction coefficient V~2(j)subscript~𝑉2𝑗\tilde{V}_{2}(j)over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_j ) is no longer an operator but an order-dependent number
V~2(j)=2m*d(1+cos(π(j+I)))+g2.subscript~𝑉2𝑗superscriptPlanck-constant-over-2-pi2superscript𝑚𝑑1𝜋𝑗𝐼𝑔2\tilde{V}_{2}(j)=\frac{\hbar^{2}}{m^{*}d}(1+\cos(\pi(j+I)))+\frac{g}{2}.over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_j ) = divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_d end_ARG ( 1 + roman_cos ( italic_π ( italic_j + italic_I ) ) ) + divide start_ARG italic_g end_ARG start_ARG 2 end_ARG . (41b)

The specific form chosen for the coefficient V~2(j)subscript~𝑉2𝑗\tilde{V}_{2}(j)over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_j ) is somewhat arbitrary; any function that alternates between 00 and 2222 depending on the particle order index j𝑗jitalic_j would work for the expression in parentheses. In another ordering sector 𝒳p1p2pN1subscript𝒳subscript𝑝1subscript𝑝2subscript𝑝𝑁1\mathcal{X}_{p_{1}p_{2}\cdots p_{N-1}}caligraphic_X start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⋯ italic_p start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, where the labelled particles are in the order xp1xp2xpNsubscript𝑥subscript𝑝1subscript𝑥subscript𝑝2subscript𝑥subscript𝑝𝑁x_{p_{1}}\leq x_{p_{2}}\leq\cdots\leq x_{p_{N}}italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≤ italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≤ ⋯ ≤ italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_POSTSUBSCRIPT, the only change to Hcont.subscript𝐻contH_{\mathrm{cont.}}italic_H start_POSTSUBSCRIPT roman_cont . end_POSTSUBSCRIPT is that the delta function in Eq. 41a becomes δ(xpjxpj+1)𝛿subscript𝑥subscript𝑝𝑗subscript𝑥subscript𝑝𝑗1\delta(x_{p_{j}}-x_{p_{j+1}})italic_δ ( italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ).

In the limit d𝑑d\to\inftyitalic_d → ∞, the coefficient V~2(j)subscript~𝑉2𝑗\tilde{V}_{2}(j)over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_j ) forces a hard-core constraint between particles that have antisymmetric exchanges (i.e. that behave like fermions to each other). The alternating hard-core two-body interactions exclude triple coincidences and so again we choose to leave out the three-body interaction term from Eq. 37 in Eq. 41a. Note that the total Hamiltonian only depends on the relative coordinates and the interactions are invariant under Galilean transformations and permutations of particle labels. Time reversal invariance is immediate from the real form of Eq. 5.2, and spatial inversion leads to the same results as for the lattice model: invariance for N𝑁Nitalic_N even and mapping between I=0𝐼0I=0italic_I = 0 and I=1𝐼1I=1italic_I = 1 for N𝑁Nitalic_N odd.

The ground state of this model can be easily calculated numerically, e.g. with a finite difference method. This allows us to compare the continuum model directly to the lattice model for traid anyons (14). We compute the ground state densities for both models and plot them in Fig. 11 (rescaling them for comparison). We see that the densities of both models match exactly, indicating that the continuum model (5.2) accurately captures the traid anyon characteristics in the low-energy, dilute limit that we observed in the lattice model.

Refer to caption
Figure 11: Comparison of the ground state densities of a system with 3 particles and traid group representation (+-+- +) for both the traid anyon lattice model (14) and its continuum limit (5.2). The lattice system contains 20202020 sites. The continuum model was solved with a finite difference method. The lattice model densities were rescaled for comparison.
Refer to caption
Figure 12: Dynamics of two identical particles on a lattice that behave like pseudofermions to each other. The position of the left and right particle are labeled by LsubscriptL\ell_{\text{L}}roman_ℓ start_POSTSUBSCRIPT L end_POSTSUBSCRIPT and RsubscriptR\ell_{\text{R}}roman_ℓ start_POSTSUBSCRIPT R end_POSTSUBSCRIPT, respectively, with LRsubscriptLsubscriptR\ell_{\text{L}}\leq\ell_{\text{R}}roman_ℓ start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ≤ roman_ℓ start_POSTSUBSCRIPT R end_POSTSUBSCRIPT (inset). The configuration space is plotted with dark (light) blue bullets corresponding to states, where both particles occupy the same (different) lattice site(s). Vertical (horizontal) lines represent tunneling processes of the left (right) particle. Thick lines correspond to tunneling matrix element to or from a state with a doubly occupied site; these are enhanced by a factor of 22\sqrt{2}square-root start_ARG 2 end_ARG. Tunneling along the red lines is associated with an extra phase of π𝜋\piitalic_π, resulting from number-dependent tunneling and giving rise to the pseudo-fermionic nature of the two particles. This leads to a flux of π𝜋\piitalic_π through the yellow-shaded plaquettes. In the continuum limit, the wave function can at most vary infinitesimally from lattice site to lattice site, because otherwise the kinetic energy would diverge. Thus, tunneling processes to a dark blue state with two particles at the same site will destructively interfere. In this way, two-particle coincidences are suppressed.

5.3 Origin of emergent hard-core interactions

The appearance of the delta-function-type contact interactions of infinite strength between two particles that behave like pseudofermions to each other can be understood intuitively from the lattice model. Let us consider a reduced model for two particles only, which locally captures the state space of the N𝑁Nitalic_N-particle traid-anyon-Hubbard model close to a coincidence of two particles that are pseudofermions to each other. This situation is plotted in Fig. 12 (see caption for a description) and also captures the pseudofermion limit θ=π𝜃𝜋\theta=\piitalic_θ = italic_π of the braid-anyon Hubbard model. The destructive interference prohibiting two particles to occupy the same site when approaching the continuum limit is associated with the appearance of plaquettes with π𝜋\piitalic_π-flux at the edge of configuration space. These are indicated by the yellow shading.

In the continuum limit, the wave function can vary at most infinitesimally from site to site, because otherwise the kinetic energy would diverge. Therefore, tunneling processes that would create a doubly occupied site (dark blue bullets) are prohibited by destructive interference, as a result of the fact that tunneling matrix elements represented by red and black lines have opposite sign. In the continuum limit two-particle coincidences are, thus, suppressed completely. This effect is taken care of by the delta-function interaction of infinite strength. Note that for two particles that behave like bosons to each other, tunneling matrix elements described by red lines have the same sign as all the others, so that no plaquette flux is induced, this effect vanishes and no interactions appear in the continuum limit.

5.4 General abelian representations for traid anyons

The derivation of the continuum limit can be generalized to arbitrary abelian representations of the traid group. As discussed in section 3 for the traid anyon lattice model, other representations are realized by constructing more complicated polynomials of the operator Njsubscript𝑁𝑗N_{j}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT (15) in the number dependent hopping phase: eiπf(Nj)nj+1superscript𝑒𝑖𝜋𝑓subscript𝑁𝑗subscript𝑛𝑗1e^{i\pi f(N_{j})n_{j+1}}italic_e start_POSTSUPERSCRIPT italic_i italic_π italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) italic_n start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. Analogous to the case of alternating exchange phases, we can then preserve the symmetry under the transformation f(Nj)f(Nj)+2𝑓subscript𝑁𝑗𝑓subscript𝑁𝑗2f(N_{j})\to f(N_{j})+2italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) → italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) + 2 by normal ordering the other operators in the exponential before taking the limit. The resulting Hamiltonian has the same structure as Eq. 37, with the new interaction coefficients

V2(x)subscript𝑉2𝑥\displaystyle V_{2}(x)italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x ) =\displaystyle== 2m*d(1cos[πf(N(x))])+g2,superscriptPlanck-constant-over-2-pi2superscript𝑚𝑑1𝜋𝑓𝑁𝑥𝑔2\displaystyle\frac{\hbar^{2}}{m^{*}d}\big{(}1-\cos[\pi f(N(x))]\big{)}+\frac{g% }{2},divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_d end_ARG ( 1 - roman_cos [ italic_π italic_f ( italic_N ( italic_x ) ) ] ) + divide start_ARG italic_g end_ARG start_ARG 2 end_ARG , (42a)
V3(x)subscript𝑉3𝑥\displaystyle V_{3}(x)italic_V start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_x ) =\displaystyle== 2m*(cos[πf(N(x))]1).superscriptPlanck-constant-over-2-pi2superscript𝑚𝜋𝑓𝑁𝑥1\displaystyle\frac{\hbar^{2}}{m^{*}}\big{(}\cos[\pi f(N(x))]-1\big{)}.divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG ( roman_cos [ italic_π italic_f ( italic_N ( italic_x ) ) ] - 1 ) . (42b)

The polynomial f(N)𝑓𝑁f(N)italic_f ( italic_N ) is constructed in the lattice model so that it takes on even integer values if a pair of particles exchanges like bosons and odd integer values if a pair of particles exchanges like fermions. Therefore, analogous to the case of alternating exchange phases, bosonic pairs do not interact in the continuum limit, whereas pairs of particles behaving like fermions experience hard-core repulsion. Alternatively, the first quantization equivalent of (42a) for the general abelian traid representation τ=(τ1τN1)𝜏subscript𝜏1subscript𝜏𝑁1\tau=(\tau_{1}\cdots\tau_{N-1})italic_τ = ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⋯ italic_τ start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT ) also could be expressed in the ordering sector 𝒳12Nsubscript𝒳12𝑁{\mathcal{X}}_{12\cdots N}caligraphic_X start_POSTSUBSCRIPT 12 ⋯ italic_N end_POSTSUBSCRIPT as

V~2(j)=2m*d(1τj)+g2.subscript~𝑉2𝑗superscriptPlanck-constant-over-2-pi2superscript𝑚𝑑1subscript𝜏𝑗𝑔2\tilde{V}_{2}(j)=\frac{\hbar^{2}}{m^{*}d}(1-\tau_{j})+\frac{g}{2}.over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_j ) = divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_d end_ARG ( 1 - italic_τ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) + divide start_ARG italic_g end_ARG start_ARG 2 end_ARG . (43)

For abelian representations of the traid group, with more than two phases τi=+1subscript𝜏𝑖1\tau_{i}=+1italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = + 1 next to each other, i.e. τi=τi+1=+1subscript𝜏𝑖subscript𝜏𝑖11\tau_{i}=\tau_{i+1}=+1italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT = + 1 or (++)(\cdots++\cdots)( ⋯ + + ⋯ ), one finds sequences of three or more neighboring particles that behave like bosons to each other. In these cases, the continuum limit of our traid-anyon-Hubbard model does not possess emergent three-body hard-core interactions within these bosonic sequences, unless we include such a three-body hard-core constraint by hand. A proper continuum limit is approached even if we do not enforce three-body hard-core interactions on the level of the lattice model.

This raises the interesting question of whether the definition of traid anyons in the continuum actually requires three-body hard-core interactions between all particles, or whether they are required only among those triples of neighboring particles that do not all behave like bosons to each other. The continuum limit of the interaction coefficient V3(x)subscript𝑉3𝑥V_{3}(x)italic_V start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_x ) (42b), which is not hard-core, suggests that that the latter is true. An alternative approach is to include a three-body hard-core constraint on the level of the tight-binding Hamiltonian as has been done also here and previously also in Refs. [36, 37]. This leads to a continuum limit with a three-body hard-core constraint among all triplets that is consistent with the traid-anyon wave functions constructed previously [16, 17]. The differences between these approaches is a subject for future investigation.

5.5 Mapping between bosonic and traid anyon wave functions

Refer to caption
Figure 13: The relative configuration space for three particles in a coordinate system (5.5). The configuration space is divided into six sectors depending on the order of the particles, indicated by inequalities of the form xp1<xp2<xp3subscript𝑥subscript𝑝1subscript𝑥subscript𝑝2subscript𝑥subscript𝑝3x_{p_{1}}<x_{p_{2}}<x_{p_{3}}italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT < italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT < italic_x start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_POSTSUBSCRIPT. Two-body coincidences xixj=0subscript𝑥𝑖subscript𝑥𝑗0x_{i}-x_{j}=0italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 0 occur at the colored, dashed lines separating the sectors at φ{2π/3,π/3,0,π/3,2π/3}𝜑2𝜋3𝜋30𝜋32𝜋3\varphi\in\{-2\pi/3,-\pi/3,0,\pi/3,2\pi/3\}italic_φ ∈ { - 2 italic_π / 3 , - italic_π / 3 , 0 , italic_π / 3 , 2 italic_π / 3 }, where tanφ=z2/z1𝜑subscript𝑧2subscript𝑧1\tan\varphi=z_{2}/z_{1}roman_tan italic_φ = italic_z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. The dashed red half-lines at φ=0𝜑0\varphi=0italic_φ = 0, 2π/32𝜋32\pi/32 italic_π / 3, and 2π/32𝜋3-2\pi/3- 2 italic_π / 3 correspond to configurations where the leftmost and center particles coincide. The green half-lines at φ=π/3𝜑𝜋3\varphi=\pi/3italic_φ = italic_π / 3, π𝜋\piitalic_π, and 2π/32𝜋3-2\pi/3- 2 italic_π / 3 correspond to configurations where the center and rightmost particles coincide. A hard-core three-body constraint excludes the origin of this coordinate system.

In this last subsection on the continuum limit, we discuss the mapping of bosonic states that evolve according to the Hamiltonian (5.2) and anyonic states that obey the symmetries given by the traid group [16, 17]. For simplicity, we will consider the case of three particles with g=0𝑔0g=0italic_g = 0 as the simplest non-trivial scenario. As shown in Fig. 13, the structure of interactions is revealed by transforming from the particle coordinates 𝐱=(x1,x2,x3)3𝐱subscript𝑥1subscript𝑥2subscript𝑥3superscript3{\bf x}=(x_{1},x_{2},x_{3})\in{\mathbb{R}}^{3}bold_x = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) ∈ blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT to Jacobi coordinates 𝐳=(z0,z1,z2)𝐳subscript𝑧0subscript𝑧1subscript𝑧2{\bf z}=(z_{0},z_{1},z_{2})bold_z = ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ):

z0subscript𝑧0\displaystyle z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =\displaystyle== 13(x1+x2+x3),13subscript𝑥1subscript𝑥2subscript𝑥3\displaystyle\frac{1}{\sqrt{3}}\left(x_{1}+x_{2}+x_{3}\right),divide start_ARG 1 end_ARG start_ARG square-root start_ARG 3 end_ARG end_ARG ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) ,
z1subscript𝑧1\displaystyle z_{1}italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =\displaystyle== 16(2x3x1x2),162subscript𝑥3subscript𝑥1subscript𝑥2\displaystyle\frac{1}{\sqrt{6}}(2x_{3}-x_{1}-x_{2}),divide start_ARG 1 end_ARG start_ARG square-root start_ARG 6 end_ARG end_ARG ( 2 italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ,
z2subscript𝑧2\displaystyle z_{2}italic_z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =\displaystyle== 12(x2x1),12subscript𝑥2subscript𝑥1\displaystyle\frac{1}{\sqrt{2}}(x_{2}-x_{1}),divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG ( italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) , (44)

where z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is proportional to the center-of-mass, z2subscript𝑧2z_{2}italic_z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT proportional to the relative position of the particles labelled 1111 and 2222, and z1subscript𝑧1z_{1}italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT proportional to the relative position of particle 3333 with the center-of-mass of particles 1111 and 2222, and all of these have been normalized so that the transformation 𝐱𝐳𝐱𝐳{\bf x}\to{\bf z}bold_x → bold_z is orthogonal and j=132/xj2=j=022/zj2superscriptsubscript𝑗13superscript2superscriptsubscript𝑥𝑗2superscriptsubscript𝑗02superscript2superscriptsubscript𝑧𝑗2\sum_{j=1}^{3}\partial^{2}/\partial x_{j}^{2}=\sum_{j=0}^{2}\partial^{2}/% \partial z_{j}^{2}∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ∂ italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Converting to polar coordinates in the relative plane,

tanφ=z2z1,ρ=z12+z22,formulae-sequence𝜑subscript𝑧2subscript𝑧1𝜌superscriptsubscript𝑧12superscriptsubscript𝑧22\displaystyle\tan\varphi=\frac{z_{2}}{z_{1}},\rho=\sqrt{z_{1}^{2}+z_{2}^{2}},roman_tan italic_φ = divide start_ARG italic_z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG , italic_ρ = square-root start_ARG italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (45)

gives the relative radius ρ𝜌\rhoitalic_ρ, describing the size of the three-body configuration, and the angular coordinate φ𝜑\varphiitalic_φ describing the relative shape [38].

The relative coordinate plane is depicted in Fig. 13 and the symmetry revealed there explains why Jacobi coordinates are convenient. As φ𝜑\varphiitalic_φ is varied from π𝜋-\pi- italic_π to π𝜋\piitalic_π for fixed ρ𝜌\rhoitalic_ρ, the three particles execute a cyclic exchange through six two-body coincidences. Three angles denoted α1subscript𝛼1\alpha_{1}italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT correspond to coincidences of the first two particles (xi=xj<xksubscript𝑥𝑖subscript𝑥𝑗subscript𝑥𝑘x_{i}=x_{j}<x_{k}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT < italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT),

α1{2π3,0,2π3},subscript𝛼12𝜋302𝜋3\alpha_{1}\in\left\{-\frac{2\pi}{3},0,\frac{2\pi}{3}\right\},italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∈ { - divide start_ARG 2 italic_π end_ARG start_ARG 3 end_ARG , 0 , divide start_ARG 2 italic_π end_ARG start_ARG 3 end_ARG } , (46a)
and three angles denoted α2subscript𝛼2\alpha_{2}italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT correspond to coincidences of the second two particles (xk<xi=xjsubscript𝑥𝑘subscript𝑥𝑖subscript𝑥𝑗x_{k}<x_{i}=x_{j}italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT < italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT),
α2{π3,π3,π},subscript𝛼2𝜋3𝜋3𝜋\alpha_{2}\in\left\{-\frac{\pi}{3},\frac{\pi}{3},\pi\right\},italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ∈ { - divide start_ARG italic_π end_ARG start_ARG 3 end_ARG , divide start_ARG italic_π end_ARG start_ARG 3 end_ARG , italic_π } , (46b)

with ijk𝑖𝑗𝑘i\neq j\neq kitalic_i ≠ italic_j ≠ italic_k. Expressed in Jacobi cylindrical coordinates, any bosonic wave function Ψ(z0,ρ,φ)Ψsubscript𝑧0𝜌𝜑\Psi(z_{0},\rho,\varphi)roman_Ψ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_φ ) must satisfy the following boundary conditions at both α1subscript𝛼1\alpha_{1}italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and α2subscript𝛼2\alpha_{2}italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT to be symmetric under any two-particle exchange:

limϵ0Ψ(z0,ρ,αi+ϵ)subscriptitalic-ϵ0Ψsubscript𝑧0𝜌subscript𝛼𝑖italic-ϵ\displaystyle\lim_{\epsilon\to 0}\Psi(z_{0},\rho,\alpha_{i}+\epsilon)roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT roman_Ψ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ϵ ) =\displaystyle== limϵ0Ψ(z0,ρ,αiϵ),subscriptitalic-ϵ0Ψsubscript𝑧0𝜌subscript𝛼𝑖italic-ϵ\displaystyle\lim_{\epsilon\to 0}\Psi(z_{0},\rho,\alpha_{i}-\epsilon),roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT roman_Ψ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_ϵ ) , (47a)
limϵ0Ψ(z0,ρ,αi+ϵ)subscriptitalic-ϵ0superscriptΨsubscript𝑧0𝜌subscript𝛼𝑖italic-ϵ\displaystyle\lim_{\epsilon\to 0}\Psi^{\prime}(z_{0},\rho,\alpha_{i}+\epsilon)roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT roman_Ψ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ϵ ) =\displaystyle== limϵ0Ψ(z0,ρ,αiϵ),subscriptitalic-ϵ0superscriptΨsubscript𝑧0𝜌subscript𝛼𝑖italic-ϵ\displaystyle-\lim_{\epsilon\to 0}\Psi^{\prime}(z_{0},\rho,\alpha_{i}-\epsilon),- roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT roman_Ψ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_ϵ ) , (47b)

where Ψ(z0,ρ,φ)=Ψ/φsuperscriptΨsubscript𝑧0𝜌𝜑Ψ𝜑\Psi^{\prime}(z_{0},\rho,\varphi)=\partial\Psi/\partial\varphiroman_Ψ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_φ ) = ∂ roman_Ψ / ∂ italic_φ and the extra minus sign in Eq. 47b accounts for the reflection of φ𝜑\varphiitalic_φ in the derivative.

In the Jacobi relative coordinates, the Hamiltonian (5.2) for the sector x1x2x3subscript𝑥1subscript𝑥2subscript𝑥3x_{1}\leq x_{2}\leq x_{3}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT has the form

Hcont.subscript𝐻cont\displaystyle H_{\mathrm{cont.}}italic_H start_POSTSUBSCRIPT roman_cont . end_POSTSUBSCRIPT =\displaystyle== j=0222m*2zj2+(1τ1)22m*dδ(φ)ρ+(1τ2)22m*dδ(φπ/3)ρ,superscriptsubscript𝑗02superscriptPlanck-constant-over-2-pi22superscript𝑚superscript2superscriptsubscript𝑧𝑗21subscript𝜏1superscriptPlanck-constant-over-2-pi22superscript𝑚𝑑𝛿𝜑𝜌1subscript𝜏2superscriptPlanck-constant-over-2-pi22superscript𝑚𝑑𝛿𝜑𝜋3𝜌\displaystyle-\sum_{j=0}^{2}\frac{\hbar^{2}}{2m^{*}}\frac{\partial^{2}}{% \partial z_{j}^{2}}+(1-\tau_{1})\frac{\hbar^{2}}{\sqrt{2}m^{*}d}\frac{\delta(% \varphi)}{\rho}+(1-\tau_{2})\frac{\hbar^{2}}{\sqrt{2}m^{*}d}\frac{\delta(% \varphi-\pi/3)}{\rho},- ∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + ( 1 - italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG square-root start_ARG 2 end_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_d end_ARG divide start_ARG italic_δ ( italic_φ ) end_ARG start_ARG italic_ρ end_ARG + ( 1 - italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG square-root start_ARG 2 end_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_d end_ARG divide start_ARG italic_δ ( italic_φ - italic_π / 3 ) end_ARG start_ARG italic_ρ end_ARG , (48)

and symmetric forms in other sectors 444Note that in the intrinsic approach to topological exchange statistics (see App. C), indistinguishability effectively restricts the configuration space to only one of the six sectors depicted in Fig. 13.. For traid representations with τ1=1subscript𝜏11\tau_{1}=-1italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - 1, the red dashed half-lines at α1subscript𝛼1\alpha_{1}italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT in Fig. 13 represent infinite-strength, contact interactions between the first and second particle in the continuum limit d0𝑑0d\to 0italic_d → 0. This forces an additional condition on the wave function

Ψ(z0,ρ,α1)=0forτ1=1.formulae-sequenceΨsubscript𝑧0𝜌subscript𝛼10forsubscript𝜏11\Psi(z_{0},\rho,\alpha_{1})=0\quad\text{for}\quad\tau_{1}=-1.roman_Ψ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = 0 for italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - 1 . (49a)
Similarly, for traid representations with τ2=1subscript𝜏21\tau_{2}=-1italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - 1, the green half-lines at α2subscript𝛼2\alpha_{2}italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT represent infinite-strength delta functions between the second and third particle and this forces the condition
Ψ(z0,ρ,α2)=0forτ2=1.formulae-sequenceΨsubscript𝑧0𝜌subscript𝛼20forsubscript𝜏21\Psi(z_{0},\rho,\alpha_{2})=0\quad\text{for}\quad\tau_{2}=-1.roman_Ψ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = 0 for italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - 1 . (49b)

Because the many-body wave function vanishes at certain coincidences, it costs no extra energy to flip the sign of the wave function on one side of such a coincidence. This allows us to construct maps from bosonic wave functions of Eq. 48 to fermionic or traid-anyonic wave functions Ψ~(z0,ρ,φ)~Ψsubscript𝑧0𝜌𝜑\tilde{\Psi}(z_{0},\rho,\varphi)over~ start_ARG roman_Ψ end_ARG ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_φ ) that are antisymmetric with respect to the exchange of the two leftmost particles, if τ1=1subscript𝜏11\tau_{1}=-1italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - 1, and/or the two rightmost particles, if τ2=1subscript𝜏21\tau_{2}=-1italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - 1:

limϵ0Ψ~(z0,ρ,αi+ϵ)subscriptitalic-ϵ0~Ψsubscript𝑧0𝜌subscript𝛼𝑖italic-ϵ\displaystyle\lim_{\epsilon\to 0}\tilde{\Psi}(z_{0},\rho,\alpha_{i}+\epsilon)roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT over~ start_ARG roman_Ψ end_ARG ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ϵ ) =\displaystyle== τilimϵ0Ψ~(z0,ρ,αiϵ)subscript𝜏𝑖subscriptitalic-ϵ0~Ψsubscript𝑧0𝜌subscript𝛼𝑖italic-ϵ\displaystyle\tau_{i}\lim_{\epsilon\to 0}\tilde{\Psi}(z_{0},\rho,\alpha_{i}-\epsilon)italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT over~ start_ARG roman_Ψ end_ARG ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_ϵ ) (50a)
limϵ0Ψ~(z0,ρ,αi+ϵ)subscriptitalic-ϵ0superscript~Ψsubscript𝑧0𝜌subscript𝛼𝑖italic-ϵ\displaystyle\lim_{\epsilon\to 0}\tilde{\Psi}^{\prime}(z_{0},\rho,\alpha_{i}+\epsilon)roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ϵ ) =\displaystyle== τilimϵ0Ψ~(z0,ρ,αiϵ)..subscript𝜏𝑖subscriptitalic-ϵ0superscript~Ψsubscript𝑧0𝜌subscript𝛼𝑖italic-ϵ\displaystyle-\tau_{i}\lim_{\epsilon\to 0}\tilde{\Psi}^{\prime}(z_{0},\rho,% \alpha_{i}-\epsilon)..- italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_ϵ ) . . (50b)

In the fermionic case ()(--)( - - ), the analogy to the Tonks-Girardeau gas is evident; infinite-strength two-body interactions mimic the exclusion provided by fermion statistics. In contrast, for the anyonic representations (+)(+-)( + - ) and (+)(-+)( - + ), the two-body hard-core interactions depend on the relative location of the third particle. We will see below that, as a result, the mapping from bosonic to traid-anyon wave functions is not one-to-one.

To make this mapping more explicit, consider the special case when there is either no trapping potential or a harmonic trapping potential and there are no additional two-body interactions. Then, we can find exact solutions by recognizing that in the d0𝑑0d\to 0italic_d → 0 limit the Hamiltonian (48) separates in z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, ρ𝜌\rhoitalic_ρ, and φ𝜑\varphiitalic_φ coordinates. Call F(φ)𝐹𝜑F(\varphi)italic_F ( italic_φ ) the angular factor of the three-body wave function, i.e. Ψ(z0,ρ,φ)=χ(z0,ρ)F(φ)Ψsubscript𝑧0𝜌𝜑𝜒subscript𝑧0𝜌𝐹𝜑\Psi(z_{0},\rho,\varphi)=\chi(z_{0},\rho)F(\varphi)roman_Ψ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ , italic_φ ) = italic_χ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ ) italic_F ( italic_φ ). In between coincidences, there is no angular potential, so the angular wave function has the free form F(φ)=Aexp(iμφ)+Bexp(iμφ)𝐹𝜑𝐴𝑖𝜇𝜑𝐵𝑖𝜇𝜑F(\varphi)=A\exp(i\mu\varphi)+B\exp(-i\mu\varphi)italic_F ( italic_φ ) = italic_A roman_exp ( italic_i italic_μ italic_φ ) + italic_B roman_exp ( - italic_i italic_μ italic_φ ) for μ0𝜇0\mu\geq 0italic_μ ≥ 0. Combining the bosonic requirement (5.5) with the representation-dependent two-body coincidence conditions (5.5), the lowest energy, unnormalized solutions F(τ1,τ2)(φ)subscript𝐹subscript𝜏1subscript𝜏2𝜑F_{(\tau_{1},\tau_{2})}(\varphi)italic_F start_POSTSUBSCRIPT ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT ( italic_φ ) for the representations (τ1,τ2)subscript𝜏1subscript𝜏2(\tau_{1},\tau_{2})( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) are

F(++)(φ)subscript𝐹absent𝜑\displaystyle F_{(++)}(\varphi)italic_F start_POSTSUBSCRIPT ( + + ) end_POSTSUBSCRIPT ( italic_φ ) =\displaystyle== 1,1\displaystyle 1,1 , (51a)
F()(φ)subscript𝐹absent𝜑\displaystyle F_{(--)}(\varphi)italic_F start_POSTSUBSCRIPT ( - - ) end_POSTSUBSCRIPT ( italic_φ ) =\displaystyle== |sin(3φ)|,3𝜑\displaystyle|\sin(3\varphi)|,| roman_sin ( 3 italic_φ ) | , (51b)
F(+)(φ)subscript𝐹absent𝜑\displaystyle F_{(+-)}(\varphi)italic_F start_POSTSUBSCRIPT ( + - ) end_POSTSUBSCRIPT ( italic_φ ) =\displaystyle== |cos(3φ/2)|,3𝜑2\displaystyle|\cos(3\varphi/2)|,| roman_cos ( 3 italic_φ / 2 ) | , (51c)
F(+)(φ)subscript𝐹absent𝜑\displaystyle F_{(-+)}(\varphi)italic_F start_POSTSUBSCRIPT ( - + ) end_POSTSUBSCRIPT ( italic_φ ) =\displaystyle== |sin(3φ/2)|,3𝜑2\displaystyle|\sin(3\varphi/2)|,| roman_sin ( 3 italic_φ / 2 ) | , (51d)

corresponding to bosons [(+,+)(+,+)( + , + )], fermions [(,)(-,-)( - , - )], and traid anyons [(+,)(+,-)( + , - ), (,+)(-,+)( - , + )]. The last three wave functions (51b-51d) are depicted by the solid lines in Fig. 14.

For the representation ()(--)( - - ), the antisymmetric boson-fermion map of Girardeau [39, 40, 41] maps the bosonic wave function (51b) depicted in Fig. 14(a) to the corresponding fermionic wave function by flipping the sign at each two-body coincidence. The resulting wave function F~()(φ)=sin(3φ)subscript~𝐹absent𝜑3𝜑\tilde{F}_{(--)}(\varphi)=\sin(3\varphi)over~ start_ARG italic_F end_ARG start_POSTSUBSCRIPT ( - - ) end_POSTSUBSCRIPT ( italic_φ ) = roman_sin ( 3 italic_φ ), depicted by a dashed line in Fig. 14(a), has the symmetries under exchange of particle labels that we expect for fermions: F~(αi+ϵ)=F~(αiϵ)~𝐹subscript𝛼𝑖italic-ϵ~𝐹subscript𝛼𝑖italic-ϵ\tilde{F}(\alpha_{i}+\epsilon)=-\tilde{F}(\alpha_{i}-\epsilon)over~ start_ARG italic_F end_ARG ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ϵ ) = - over~ start_ARG italic_F end_ARG ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_ϵ ) and F~(αi+ϵ)=F~(αiϵ)superscript~𝐹subscript𝛼𝑖italic-ϵsuperscript~𝐹subscript𝛼𝑖italic-ϵ\tilde{F}^{\prime}(\alpha_{i}+\epsilon)=\tilde{F}^{\prime}(\alpha_{i}-\epsilon)over~ start_ARG italic_F end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ϵ ) = over~ start_ARG italic_F end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_ϵ ) for ϵ0italic-ϵ0\epsilon\to 0italic_ϵ → 0. Note that it is possible to satisfy these requirements with single-valued functions on the interval φ(π,π]𝜑𝜋𝜋\varphi\in(-\pi,\pi]italic_φ ∈ ( - italic_π , italic_π ].

In contrast, when we attempt to construct a similar map from the bosonic to the anyonic wave functions for the (+)(+-)( + - ) and (+)(-+)( - + ) representations, we find the function that satisfies (5.5) must be double-valued. For these representations, traid anyon exchange statistics impose the symmetries at α1subscript𝛼1\alpha_{1}italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and α2subscript𝛼2\alpha_{2}italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT: F~(αi+ϵ)=τiF~(αiϵ)~𝐹subscript𝛼𝑖italic-ϵsubscript𝜏𝑖~𝐹subscript𝛼𝑖italic-ϵ\tilde{F}(\alpha_{i}+\epsilon)=\tau_{i}\tilde{F}(\alpha_{i}-\epsilon)over~ start_ARG italic_F end_ARG ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ϵ ) = italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over~ start_ARG italic_F end_ARG ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_ϵ ) and F~(αi+ϵ)=τiF~(αiϵ)superscript~𝐹subscript𝛼𝑖italic-ϵsubscript𝜏𝑖superscript~𝐹subscript𝛼𝑖italic-ϵ\tilde{F}^{\prime}(\alpha_{i}+\epsilon)=-\tau_{i}\tilde{F}^{\prime}(\alpha_{i}% -\epsilon)over~ start_ARG italic_F end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ϵ ) = - italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over~ start_ARG italic_F end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_ϵ ) for ϵ0italic-ϵ0\epsilon\to 0italic_ϵ → 0. Unlike the fermionic case, flipping the sign of the wave functions at the hard-core two-body coincidences no longer yields a single-valued map; the sign of the resulting function depends on where we begin the sequence of sign-flipping exchanges. For example, consider the function F(+)subscript𝐹absentF_{(+-)}italic_F start_POSTSUBSCRIPT ( + - ) end_POSTSUBSCRIPT (51c). One possibility for flipping the signs is depicted by the dashed line in Fig. 14(b), where F(+)subscript𝐹absentF_{(+-)}italic_F start_POSTSUBSCRIPT ( + - ) end_POSTSUBSCRIPT has been sign-flipped at the two-body hard-core coincidences at φ=π/3𝜑𝜋3\varphi=-\pi/3italic_φ = - italic_π / 3 and π/3𝜋3\pi/3italic_π / 3, giving the function F~(+)(φ)=cos(3φ/2)subscript~𝐹absent𝜑3𝜑2\tilde{F}_{(+-)}(\varphi)=\cos(3\varphi/2)over~ start_ARG italic_F end_ARG start_POSTSUBSCRIPT ( + - ) end_POSTSUBSCRIPT ( italic_φ ) = roman_cos ( 3 italic_φ / 2 ). However, choosing to start in the domain φ[π,π/3]𝜑𝜋𝜋3\varphi\in[-\pi,-\pi/3]italic_φ ∈ [ - italic_π , - italic_π / 3 ] and antisymmetrizing across φ=π𝜑𝜋\varphi=-\piitalic_φ = - italic_π and π/3𝜋3-\pi/3- italic_π / 3 gives the opposite sign F~(+)(φ)=cos(3φ/2)subscript~𝐹absent𝜑3𝜑2\tilde{F}_{(+-)}(\varphi)=-\cos(3\varphi/2)over~ start_ARG italic_F end_ARG start_POSTSUBSCRIPT ( + - ) end_POSTSUBSCRIPT ( italic_φ ) = - roman_cos ( 3 italic_φ / 2 ). This indicates that to implement the correct exchange statistics, the traid anyon wave function for the (+)(+-)( + - ) representation must be double-valued functions on the interval φ(π,π]𝜑𝜋𝜋\varphi\in(-\pi,\pi]italic_φ ∈ ( - italic_π , italic_π ].

Note that the three-body wavefunctions F~(φ)~𝐹𝜑\tilde{F}(\varphi)over~ start_ARG italic_F end_ARG ( italic_φ ) that we have obtained here by taking the continuum limit of our lattice model, directly correspond to the continuum traid-anyon wavefunctions constructed previously in Ref. [16] from general properties of the traid group. This provides another justification of the intuitive construction of the lattice model presented here.

In the presence of non-quadratic trapping potentials or additional interactions between the particles, the above reasoning remains valid, except that separability is lost and the wave functions do not have a simple form like (5.5). We can also generalize Eqs. (5.5-5.5) to the case of more than three particles by requiring corresponding symmetries dictated by τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for each of the relative coordinates of neighboring particles, while keeping all other coordinates fixed. This mapping between bosons and traid anyons provides traid-anyon wavfunctions in the continuum and corresponds to the generalized Jordan-Wigner transformation that we employed for the traid-anyon-Hubbard model.

Refer to caption
Figure 14: This figure depicts the angular factor F(τ1τ2)(φ)subscript𝐹subscript𝜏1subscript𝜏2𝜑F_{(\tau_{1}\tau_{2})}(\varphi)italic_F start_POSTSUBSCRIPT ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT ( italic_φ ) of bosonic three-body wave functions solving the Hamiltonian (48) for the abelian traid group representations ()(--)( - - ), (+)(+-)( + - ), and (+)(-+)( - + ), respectively. The solid lines correspond to the lowest energy solutions. The thicker dotted lines show corresponding traid-anyon wave functions F~(τ1τ2)(φ)subscript~𝐹subscript𝜏1subscript𝜏2𝜑\tilde{F}_{(\tau_{1}\tau_{2})}(\varphi)over~ start_ARG italic_F end_ARG start_POSTSUBSCRIPT ( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT ( italic_φ ). In the pseudofermion case depicted in subfigure (a), infinite-strength contact interactions force nodes at every two-body coincidence for the black, solid wave function. The gray, thicker dotted line represents the angular wave function of the lowest energy free fermion state. This function has the same energy as its bosonic counterpart and is single-valued on the interval φ(π,π]𝜑𝜋𝜋\varphi\in(-\pi,\pi]italic_φ ∈ ( - italic_π , italic_π ]. It can be related to the bosonic wave function by the totally antisymmetric boson-fermion map of Girardeau [39].The solid lines in subfigures (b) and (c) depict the bosonic wave functions correspond to traid anyon representations (+)(+-)( + - ) and (+)(-+)( - + ), respectively. These have nodes at alternating two-body coincidences. The dotted lines represent one sheet of the corresponding multi-valued anyonic wave functions.

6 Conclusions

The main result of our paper is the construction of a lattice model for abelian traid anyons in one dimension. The starting point of this construction is the observation that the discreteness of the lattice’s configuration space (which is given by Fock states with sharp site occupation numbers) allows to associate the exchange of two particles on neighboring sites with different paths in that space. The exchange can either be accomplished by the left particle hopping to the right onto the coincidence followed by a second hop to the right or by the right particle hopping to the left onto the coincidence followed by a second hop the left. One can argue that distinguishing both processes, and associating them with the left or the right particle passing “in front” of the other one, allows to define the braiding of particles in one dimension and that this physics is realized by the (braid) anyon Hubbard model [11]. The phase picked up by the wave function under such particle exchange processes directly corresponds to a geometric phase around (or flux through) closed loops in configuration space. Abelian representations for braid anyons are characterized by a single angle θ𝜃\thetaitalic_θ that determines the exchange phase and smoothly interpolates between bosons (θ=0𝜃0\theta=0italic_θ = 0) and fermions (θ=π𝜃𝜋\theta=\piitalic_θ = italic_π). For abelian traid anyons, in turn, the phase θisubscript𝜃𝑖\theta_{i}italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT associated with the exchange of the i𝑖iitalic_ith and (i+1)𝑖1(i+1)( italic_i + 1 )st particle from the left depends on the ordinality i𝑖iitalic_i, but can take two possible values, 00 or π𝜋\piitalic_π only, giving rise to a countable number of abelian representations (τ1,τ2,,τN1)subscript𝜏1subscript𝜏2subscript𝜏𝑁1(\tau_{1},\tau_{2},\dots,\tau_{N-1})( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_τ start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT ) with τi=exp(iθi)subscript𝜏𝑖𝑖subscript𝜃𝑖\tau_{i}=\exp(i\theta_{i})italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = roman_exp ( italic_i italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ).

In order to construct a lattice model exhibiting these traid anyon statistics, we first consider bosonic particles with non-local occupation-number-dependent hopping (or Peierls) phases. These are chosen to give rise to the geometric phases in Fock space corresponding to the desired abelian representation of traid anyons. This approach resembles the composite-particle description of braid anyons in two dimensions, where anyons are mapped to charged bosons (or fermions) to which a flux tube is attached that gives rise to the geometric phase θ𝜃\thetaitalic_θ when one particle is moved around another one. In a second step, we then introduce traid-anyon annihilation and creation operators, with respect to which the kinetic term of the Hamiltonian becomes quadratic and local, and derive their commutation relations. This construction provides an intuitive approach to traid anyons.

Investigating the ground-state properties of this traid-anyon-Hubbard model, we find various indications also of an approximate Haldane-type exclusion statistics. For instance, we observe generalized Friedel oscillations, nearly integer occupations of the natural orbits, and a step-wise change of the chemical potential with particle number.

Finally, we take the continuum limit of our lattice model and find a Hamiltonian, whose eigenstates indeed correspond to traid-anyon wavefunctions, as they were constructed previously from the properties of the traid group [16]. This justifies a posteriori the construction of our model. Moreover, the continuum limit of the traid-anyon Hubbard model is found to be Galilean invariant. This is in clear contrast to the continuum limit of the braid-anyon-Hubbard model, which breaks Galilean invariance by a Hamiltonian that, like the Kundu model [42, 43, 44, 45, 46, 47] and chiral BF model [48, 49, 50, 51, 52], depends on the current-density operator [15]. We attribute this difference to the fact that traid anyons are naturally defined also in the continuum in 1D using the intrinsic approach to topological exchange statistics. For braid anyons, on the other hand, this is not the case as they require to distinguish exchange processes with the left particle passing over or under the right one during an exchange, which in a 1D continuum system can only be mimicked dynamically by a velocity dependence.

Our results indicate that the possibilities for non-standard statistics might in some sense be even richer in one dimension than in two dimensions, at least when considering discrete lattice models. We have shown how 1D bosonic lattice models with density-dependent Peierls phases can implement abelian anyons that possess two completely different forms of exchange statistics, each corresponding to a different way of breaking the symmetric group: braid statistics and traid statistics. In both cases, the theoretical link between the interacting bosonic lattice Hamiltonian and the ‘free’ anyonic Hamiltonian is provided by a Jordan-Wigner-like gauge transformation that can be expressed in the form

aj=eiΘ(Nj)bj,subscript𝑎𝑗superscript𝑒𝑖Θsubscript𝑁𝑗subscript𝑏𝑗a_{j}=e^{i\Theta(N_{j})}b_{j},italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT italic_i roman_Θ ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (52)

where Θ(Nj)Θsubscript𝑁𝑗\Theta(N_{j})roman_Θ ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) is a so-called non-local string operator that depends on Njsubscript𝑁𝑗N_{j}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, the number of particles to the left of (and including) site j𝑗jitalic_j. For the braid-anyon Hubbard model, the function has the simple form Θ(Nj)=θNjΘsubscript𝑁𝑗𝜃subscript𝑁𝑗\Theta(N_{j})=\theta N_{j}roman_Θ ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = italic_θ italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and the bosonic Hamiltonian possesses local interactions but has a continuum limit that is not Galilean-invariant and may not lead to a consistent boson-anyon mapping [41]. For the traid-anyon Hubbard model, the function has a more complicated form Θ(Nj)=πf(Nj)Θsubscript𝑁𝑗𝜋𝑓subscript𝑁𝑗\Theta(N_{j})=-\pi f(N_{j})roman_Θ ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = - italic_π italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) where f(Nj)𝑓subscript𝑁𝑗f(N_{j})italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) is an integer function that depends on the traid group representation. The resulting bosonic Hamiltonian possesses non-local interactions that depend on Njsubscript𝑁𝑗N_{j}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. It has a Galilean-invariant continuum limit and a boson-anyon mapping that is multi-valued but well-defined. The precise form of f(Nj)𝑓subscript𝑁𝑗f(N_{j})italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) depends on whether we allow or exclude three-body (or more) coincidences on the lattice. Since the specific choice of f(Nj)𝑓subscript𝑁𝑗f(N_{j})italic_f ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) results in differing continuum limits, the role of the three-body constraints merits further investigation.

These two choices for Θ(Nj)Θsubscript𝑁𝑗\Theta(N_{j})roman_Θ ( italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) were inspired by analogy with anyons in the continuum, and there are certainly other possible functions to consider. Whether these alternate forms would have physically-meaningful exchange statistics, locality properties, and continuum limits is an interesting question. One can also imagine more general string operators that depend not just on the total number of particles to the left Njsubscript𝑁𝑗N_{j}italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, but on their specific occupation-number configuration. However, as noted before, in the continuum limit the ordinality operator NjN(x)subscript𝑁𝑗𝑁𝑥N_{j}\to N(x)italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT → italic_N ( italic_x ) has a special topological property: it is invariant under homeomorphisms of a line. More generally, even indistinguishable particles can be given ‘natural’ labels on a line based on their ordinality, unlike particles in higher dimensions. While the notion of ordinality on a line may seem trivial, its topological origin may explain why 1D quantum models support such a wealth of integrable and statistical structures.

The chiral BF model as well as the braid-anyon Hubbard model have recently been realized experimentally with ultracold atoms [52, 14]. The construction of the traid-anyon Hubbard model presented here constitutes an important step also towards a first experimental realization of traid anyons in the laboratory. It is an interesting task for future research to design experimentally accessible schemes for the implementation of this (or related) traid-anyon models, though such an implementation will have to address the challenge of effectively realizing the required non-local interactions. Another interesting open question concerns the possibility of finding interacting one-dimensional systems, the excitations of which behave like traid anyons.

Acknowledgements

We thank Martin Bonkhoff and Philip Johnson for useful discussions.

Funding information

This research was funded by the Deutsche Forschungsgemeinschaft (DFG) via the Research Unit FOR 2414 under project No. 277974659. Moreover, N.H. acknowledges the support from the National Science Foundation under Grant No. NSF PHY-1748958 and the Deustcher Akademisher Austauchdienst, B.W. from the ERC grant LATIS and the EOS project CHEQS, and S.N. from the Provincia Autonoma di Trento, Q@TN, and the BMBF through the project ‘MAGICApp’.

Appendix A Topological exchange statistics

Topological exchange statistics is an alternate approach to the Symmetrization Postulate for treating indistinguishable particles in first quantization. Instead of being imposed at the outset, the transformation properties of wave functions under particle exchanges are derived from the topology of configuration space. Topological exchange statistics can reproduce the results of the Symmetrization Postulate, but can also provide additional anyonic solutions in low-dimensional system with hard-core interactions.

The key observation that underlies this approach to exchange statistics is that, when the classical configuration space 𝒬𝒬{\mathcal{Q}}caligraphic_Q for a physical system is not simply-connected, i.e. not all loops can be continuously deformed to a point, then in addition to single-valued wave functions on 𝒬𝒬{\mathcal{Q}}caligraphic_Q, multi-valued wave functions may be necessary to describe the full Hilbert space of functions of 𝒬𝒬{\mathcal{Q}}caligraphic_Q. These multi-valued functions can be formulated in several different frameworks [53, 54, 55, 56, 57].

The topological approach to exchange statistics was pioneered by researchers in path integration methods [58, 59, 53]. In 1977, Leinaas and Myrheim [1] first recognized the possibility for novel exchange statistics in low-dimensional systems; Goldin et al. independently rediscovered this result a few years later using a related approach based on current algebras [2]. The topological approach to exchange statistics was extended to orbifold configuration spaces in [16, 17].

The topological approach is also called the intrinsic approach to exchange statistics because it does not introduce arbitrary particle labels for indistinguishable particles [60]. Whereas the Symmetrization Postulate quantizes the distinguishable particle configuration space 𝒳N𝒳superscript𝑁{\mathcal{X}}\subseteq{\mathcal{M}}^{N}caligraphic_X ⊆ caligraphic_M start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT, the intrinsic approach quantizes the quotient space 𝒬=𝒳/SN𝒬𝒳subscript𝑆𝑁{\mathcal{Q}}={\mathcal{X}}/S_{N}caligraphic_Q = caligraphic_X / italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT. In the quotient map from 𝒳𝒳{\mathcal{X}}caligraphic_X to 𝒬𝒬{\mathcal{Q}}caligraphic_Q, points of 𝒳𝒳{\mathcal{X}}caligraphic_X that differ by a permutation of labels are collapsed to same point in 𝒬𝒬{\mathcal{Q}}caligraphic_Q. Unless 𝒳𝒳{\mathcal{X}}caligraphic_X excludes all two-body coincidences, the quotient space 𝒬𝒬{\mathcal{Q}}caligraphic_Q is not naturally a manifold and is better described as an orbifold [60, 61, 16, 62, 17, 63]. An advantage of the intrinsic approach is that, unlike a symmetry, true indistinguishability cannot be broken; the intrinsic approach treats particle labels as a gauge structure imposed on indistinguishable particles, not as a symmetry of configuration space  [25, 64].

In the intrinsic approach, particle exchanges are concrete continuous trajectories interchanging the particles, not abstract label permutations. These exchange trajectories start and end at configurations that differ at most by particle labels and so, in the indistinguishable configuration space 𝒬𝒬{\mathcal{Q}}caligraphic_Q, they form based loops. Under continuous deformation, equivalence classes of based loops form the fundamental (or first homotopy) group π1subscript𝜋1\pi_{1}italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT on a manifold; for orbifolds, such as 𝒬𝒬{\mathcal{Q}}caligraphic_Q, this can be generalized to the orbifold fundamental group π1*superscriptsubscript𝜋1\pi_{1}^{*}italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT [65, 66]. Unitary irreducible representations of this group classify different quantizations of the configuration space, or in the case of particle exchanges, different superselection sectors.

When {\mathcal{M}}caligraphic_M is simply-connected (i.e., π1()subscript𝜋1\pi_{1}({\mathcal{M}})italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( caligraphic_M ) is trivial) and there are no hard-core interactions, then π1*(𝒬)=SNsuperscriptsubscript𝜋1𝒬subscript𝑆𝑁\pi_{1}^{*}({\mathcal{Q}})=S_{N}italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( caligraphic_Q ) = italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT for any dimdimension\dim{\mathcal{M}}roman_dim caligraphic_M and the group of possible exchanges is equivalent to the group of particle permutations. In this case, the Symmetrization Postulate holds and there are only bosons and fermions with single-valued wave functions on 𝒳𝒳{\mathcal{X}}caligraphic_X and standard exchange statistics.

Appendix B Braid group and fractional exchange statistics

In 2D, hard-core two-body interactions create topological defects that make the configuration space not simply-connected. The paths of particle exchanges can wind around these defects, meaning that multiple topologically inequivalent exchanges may lead to the same permutation. The configuration space for N𝑁Nitalic_N indistinguishable particles on a plane 2superscript2{\mathbb{R}}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is

𝒬B2NΔ2SN,subscript𝒬𝐵superscript2𝑁subscriptΔ2subscript𝑆𝑁{\mathcal{Q}}_{B}\equiv\frac{{\mathbb{R}}^{2N}-\Delta_{2}}{S_{N}},caligraphic_Q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ≡ divide start_ARG blackboard_R start_POSTSUPERSCRIPT 2 italic_N end_POSTSUPERSCRIPT - roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_ARG , (53)

where Δ2subscriptΔ2\Delta_{2}roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is the set of points in N=2Nsuperscript𝑁superscript2𝑁{\mathcal{M}}^{N}={\mathbb{R}}^{2N}caligraphic_M start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT = blackboard_R start_POSTSUPERSCRIPT 2 italic_N end_POSTSUPERSCRIPT where at least two coordinates coincide. The space 𝒬Bsubscript𝒬𝐵{\mathcal{Q}}_{B}caligraphic_Q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is a manifold; all coincidences Δ2subscriptΔ2\Delta_{2}roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT have been removed and so 𝒬Bsubscript𝒬𝐵{\mathcal{Q}}_{B}caligraphic_Q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT contains no singular orbifold points. In the simplest case of N=2𝑁2N=2italic_N = 2, the intrinsic configuration space 𝒬Bsubscript𝒬𝐵{\mathcal{Q}}_{B}caligraphic_Q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is the direct product of a plane 2superscript2{\mathbb{R}}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for the center-of-mass coordinates and a cone with opening angle π/3𝜋3\pi/3italic_π / 3 with an excluded tip for the relative coordinates [1].

The resulting fundamental group for N𝑁Nitalic_N particles is the braid group π1(𝒬B)=BNsubscript𝜋1subscript𝒬𝐵subscript𝐵𝑁\pi_{1}({\mathcal{Q}}_{B})=B_{N}italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( caligraphic_Q start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) = italic_B start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT [4]. As described in the main text, fractional exchange statistics are given by the abelian representations of the braid group (5) with phase biβi=exp(iθi)subscript𝑏𝑖subscript𝛽𝑖𝑖subscript𝜃𝑖b_{i}\to\beta_{i}=\exp(i\theta_{i})italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT → italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = roman_exp ( italic_i italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ). This exchange phase θ𝜃\thetaitalic_θ can be associated to the Aharonov-Bohm phase in the composite charged particle/flux tube model of Wilczek [67] that gave anyons their name. Fractional exchange statistics emerge for quasiparticle excitations in certain quantum liquid and quantum spin liquid models like the fractional quantum Hall effect [8] or the Haldane-Shastry spin-chain model [34, 35]. In addition to motivating the braid-anyon-Hubbard model [11], implementing fractional exchange statistics in one-dimensional systems provided the original motivation for several continuum models, including the recently experimentally-realized chiral BF model [48, 49, 50, 51, 52], the closely-related Kundu model (also called Lieb-Liniger anyons) [42, 43, 44, 45, 46, 47], the Calogero-Sutherland anyon model [68, 25, 69], and hard-core anyon models [70, 71, 72]. Note that we have only discussed abelian braid anyons; non-abelian representations of BNsubscript𝐵𝑁B_{N}italic_B start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT are carried by multi-component, multi-valued wave functions, c.f. [73, 74].

In some of this literature on fractional exchange statistics, there exists an alternate definition to (5) for fractional exchange statistics where the transformation law for wave functions Ψ(𝐱)Ψ𝐱\Psi({\bf x})roman_Ψ ( bold_x ) is defined on 𝒳𝒳{\mathcal{X}}caligraphic_X instead of 𝒬𝒬{\mathcal{Q}}caligraphic_Q. In the simplest case for two particles, this alternate choice for the transformation has the form Ψ(x1,x2)=exp(iθ)Ψ(x2,x1)Ψsubscript𝑥1subscript𝑥2𝑖𝜃Ψsubscript𝑥2subscript𝑥1\Psi(x_{1},x_{2})=\exp(i\theta)\Psi(x_{2},x_{1})roman_Ψ ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = roman_exp ( italic_i italic_θ ) roman_Ψ ( italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ), and this generalizes to more particles (c.f. [42, 72]). Iterating this form of the transformation again, we find Ψ(x1,x2)=exp(2iθ)Ψ(x1,x2)Ψsubscript𝑥1subscript𝑥22𝑖𝜃Ψsubscript𝑥1subscript𝑥2\Psi(x_{1},x_{2})=\exp(2i\theta)\Psi(x_{1},x_{2})roman_Ψ ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = roman_exp ( 2 italic_i italic_θ ) roman_Ψ ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ), showing this wave function is indeed multi-valued as one expects for an anyonic wave function. However, there is not a mechanism for implementing the inverse transformation (i.e, an ‘under’ path with phase exp(iθ)𝑖𝜃\exp(-i\theta)roman_exp ( - italic_i italic_θ ) to match the ‘over’ path). Effectively, this equates both bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and bi1superscriptsubscript𝑏𝑖1b_{i}^{-1}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT with the same permutation on 𝒳𝒳{\mathcal{X}}caligraphic_X, so this alternate transformation rule does not describe abelian representations of the braid group.

Anyonic wave functions Ψ(𝐪)Ψ𝐪\Psi({\bf q})roman_Ψ ( bold_q ) can be lifted to functions on 𝒳𝒳{\mathcal{X}}caligraphic_X, but they will be multi-valued except for the special cases of fermions θ=π𝜃𝜋\theta=\piitalic_θ = italic_π and bosons θ=0𝜃0\theta=0italic_θ = 0. Alternatively, one can work with single-valued functions on 𝒳𝒳{\mathcal{X}}caligraphic_X for arbitrary θ𝜃\thetaitalic_θ at the cost of a gauge transformation. Famously, anyons in two-dimensions with fractional exchange statistics can be transmuted into bosons or fermions with single-valued wave functions by introducing a magnetic gauge potential [8, 6].

Appendix C Traid group and abelian traid anyon exchange statistics

In analogy to the braid group, hard-core three-body interactions in 1D also introduce topological defects to configuration space that lead to a generalization of the symmetric group. The relevant intrinsic configuration space is

𝒬T=NΔ3SN,subscript𝒬𝑇superscript𝑁subscriptΔ3subscript𝑆𝑁{\mathcal{Q}}_{T}=\frac{{\mathbb{R}}^{N}-\Delta_{3}}{S_{N}},caligraphic_Q start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = divide start_ARG blackboard_R start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT - roman_Δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG start_ARG italic_S start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_ARG , (54)

where Δ3subscriptΔ3\Delta_{3}roman_Δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is the locus of points in N=Nsuperscript𝑁superscript𝑁{\mathcal{M}}^{N}={\mathbb{R}}^{N}caligraphic_M start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT = blackboard_R start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT where at least three particle coordinates coincide. The orbifold fundamental group π1*(𝒬T)=TNsuperscriptsubscript𝜋1subscript𝒬𝑇subscript𝑇𝑁\pi_{1}^{*}({\mathcal{Q}}_{T})=T_{N}italic_π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( caligraphic_Q start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ) = italic_T start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT for N3𝑁3N\geq 3italic_N ≥ 3 indistinguishable particles is the traid group. Note that 𝒬Tsubscript𝒬𝑇{\mathcal{Q}}_{T}caligraphic_Q start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT is an orbifold, not a manifold, because not all orbifold points in Δ2subscriptΔ2\Delta_{2}roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT have been excluded [16, 17].

The configuration space 𝒬Tsubscript𝒬𝑇{\mathcal{Q}}_{T}caligraphic_Q start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT for the simplest case of N=3𝑁3N=3italic_N = 3 is represented in Fig. 15. In this depiction, it looks like a wedge q1q2q3subscript𝑞1subscript𝑞2subscript𝑞3q_{1}\leq q_{2}\leq q_{3}italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ italic_q start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT with qisubscript𝑞𝑖q_{i}\in{\mathbb{R}}italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∈ blackboard_R with an opening angle of π/3𝜋3\pi/3italic_π / 3. Note that this picture is misleading at the orbifold points, as their non-trivial topology cannot be represented in this embedding. The two half-plane faces of the wedge correspond to configurations where the leftmost and rightmost pairs of particles coincide. The line of intersection of these half-planes is the excluded locus Δ3subscriptΔ3\Delta_{3}roman_Δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT.

Refer to caption
Figure 15: Subfigure (a) shows the configuration space for three indistinguishable particles on a line 𝒬T=(2Δ3)/S3subscript𝒬𝑇superscript2subscriptΔ3subscript𝑆3{\mathcal{Q}}_{T}=({\mathbb{R}}^{2}-\Delta_{3})/S_{3}caligraphic_Q start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = ( blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - roman_Δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) / italic_S start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. See text for description; the coordinates for the sector are q1q2q3subscript𝑞1subscript𝑞2subscript𝑞3q_{1}\leq q_{2}\leq q_{3}italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ italic_q start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT where qisubscript𝑞𝑖q_{i}italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the i𝑖iitalic_ith particle from the left. The yellow boundary half-plane at the bottom is the orbifold locus q1=q2subscript𝑞1subscript𝑞2q_{1}=q_{2}italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT; the cyan boundary half-plane at the top is the orbifold locus q2=q3subscript𝑞2subscript𝑞3q_{2}=q_{3}italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. The black diagonal is the three-body coincidence locus Δ3subscriptΔ3\Delta_{3}roman_Δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. Subfigure (b) shows the relative configuration space for 𝒬Tsubscript𝒬𝑇{\mathcal{Q}}_{T}caligraphic_Q start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT along with two representative exchange paths for loops t1subscript𝑡1t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT that touch the orbifold locus q1=q2subscript𝑞1subscript𝑞2q_{1}=q_{2}italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and loops t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT that touch the orbifold locus q2=q3subscript𝑞2subscript𝑞3q_{2}=q_{3}italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. In the relative wedge, the angular coordinate φ𝜑\varphiitalic_φ ranges from φ=0𝜑0\varphi=0italic_φ = 0 to φ=π/3𝜑𝜋3\varphi=\pi/3italic_φ = italic_π / 3.

For the case of N=3𝑁3N=3italic_N = 3, the orbifold fundamental group is the traid group T3subscript𝑇3T_{3}italic_T start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT generated by two pairwise exchanges t1subscript𝑡1t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. In Fig. 15, these look like paths that start at a base point in 𝒬Tsubscript𝒬𝑇{\mathcal{Q}}_{T}caligraphic_Q start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT and touch the orbifold singularities at q1=q2subscript𝑞1subscript𝑞2q_{1}=q_{2}italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and q2=q3subscript𝑞2subscript𝑞3q_{2}=q_{3}italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, respectively. Because ti2=1superscriptsubscript𝑡𝑖21t_{i}^{2}=1italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1, a loop that touches the same orbifold singularity twice can be contracted to a loop that does not touch the boundary. Therefore, all loops are either trivial, corresponding to the identity element of T3subscript𝑇3T_{3}italic_T start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, or are alternating sequences of t1subscript𝑡1t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. Aficionados of discrete groups will recognize that T3subscript𝑇3T_{3}italic_T start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is isomorphic to the infinite dihedral group Dsubscript𝐷D_{\infty}italic_D start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, which has four one-dimensional irreducible representations corresponding to the four we have already described (++)(++)( + + ), (+)(+-)( + - ), (+)(-+)( - + ), and ()(--)( - - ); c.f. Table. 1. There is also a 1-parameter family of two-dimensional irreducible representations characterized by the angle α𝛼\alphaitalic_α between the two ‘reflections’ t1subscript𝑡1t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT; these can be parameterized as

t1(1001),t2(cosαsinαsinαcosα).formulae-sequencesubscript𝑡11001subscript𝑡2𝛼𝛼𝛼𝛼t_{1}\rightarrow\left(\begin{array}[]{cc}1&0\\ 0&-1\end{array}\right),t_{2}\rightarrow\left(\begin{array}[]{cc}\cos\alpha&% \sin\alpha\\ \sin\alpha&-\cos\alpha\end{array}\right).italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT → ( start_ARRAY start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL - 1 end_CELL end_ROW end_ARRAY ) , italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT → ( start_ARRAY start_ROW start_CELL roman_cos italic_α end_CELL start_CELL roman_sin italic_α end_CELL end_ROW start_ROW start_CELL roman_sin italic_α end_CELL start_CELL - roman_cos italic_α end_CELL end_ROW end_ARRAY ) . (55)

The single two-dimensional representation of S3subscript𝑆3S_{3}italic_S start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT occurs as one of these representations of T3Dsimilar-tosubscript𝑇3subscript𝐷T_{3}\sim D_{\infty}italic_T start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ∼ italic_D start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT corresponding to the specific angle of α=2π/3𝛼2𝜋3\alpha=2\pi/3italic_α = 2 italic_π / 3.

irrep e𝑒eitalic_e t1subscript𝑡1t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT statistics
(++)(++)( + + ) 1111 +11+1+ 1 +11+1+ 1 bosons
()(--)( - - ) 1111 11-1- 1 11-1- 1 fermions
(+)(+-)( + - ) 1111 +11+1+ 1 11-1- 1 abelian traid anyon
(+)(-+)( - + ) 1111 11-1- 1 +11+1+ 1 abelian traid anyon
0<α<π0𝛼𝜋0<\alpha<\pi0 < italic_α < italic_π 2222 00 00 non-abelian traid anyons
Table 1: Character table for identity e𝑒eitalic_e and generators t1subscript𝑡1t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT of T3Dsimilar-tosubscript𝑇3subscript𝐷T_{3}\sim D_{\infty}italic_T start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ∼ italic_D start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT. The generators t1subscript𝑡1t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT describe loops in 𝒬3subscript𝒬3{\mathcal{Q}}_{3}caligraphic_Q start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT that realize particle exchanges and their representations describe possible topological exchange statistics.

Single-component wave functions Ψ(𝐪)Ψ𝐪\Psi({\bf q})roman_Ψ ( bold_q ) in the representation (τ1τ2τN1)subscript𝜏1subscript𝜏2subscript𝜏𝑁1(\tau_{1}\tau_{2}\ldots\tau_{N-1})( italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT … italic_τ start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT ) are defined on 𝐪𝒬T𝐪subscript𝒬𝑇{\bf q}\in{\mathcal{Q}}_{T}bold_q ∈ caligraphic_Q start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT and transform under an exchange path constructed from M𝑀Mitalic_M pairwise exchanges γ=tiMti2ti1TN𝛾subscript𝑡subscript𝑖𝑀subscript𝑡subscript𝑖2subscript𝑡subscript𝑖1subscript𝑇𝑁\gamma=t_{i_{M}}\cdots t_{i_{2}}t_{i_{1}}\in T_{N}italic_γ = italic_t start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⋯ italic_t start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∈ italic_T start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT like

U^γΨ(𝐪)=(τiMτi2τi1)Ψ(𝐪).subscript^𝑈𝛾Ψ𝐪subscript𝜏subscript𝑖𝑀subscript𝜏subscript𝑖2subscript𝜏subscript𝑖1Ψ𝐪\hat{U}_{\gamma}\Psi({\bf q})=(\tau_{i_{M}}\cdots\tau_{i_{2}}\tau_{i_{1}})\Psi% ({\bf q}).over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT roman_Ψ ( bold_q ) = ( italic_τ start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⋯ italic_τ start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) roman_Ψ ( bold_q ) . (56)

In the cases of bosons or fermions, this agrees with the transformation law (2) and Ψ(𝐪)Ψ𝐪\Psi({\bf q})roman_Ψ ( bold_q ) can be lifted to a single-valued function on 𝒳T=NΔ3subscript𝒳𝑇superscript𝑁subscriptΔ3{\mathcal{X}}_{T}={\mathbb{R}}^{N}-\Delta_{3}caligraphic_X start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = blackboard_R start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT - roman_Δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. However, if not all τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are the same, then the same permutation can be executed by topologically distinguishable exchange paths. Lifting the wave function for one of these abelian mixed representations to 𝒳Tsubscript𝒳𝑇{\mathcal{X}}_{T}caligraphic_X start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT so that every exchange γTN𝛾subscript𝑇𝑁\gamma\in T_{N}italic_γ ∈ italic_T start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT is represented correctly results in a multi-valued function, as described in the main text.

Appendix D Haldane exclusion statistics

Above we have discussed topological exchange statistics, where indistinguishable particles are characterized by the transformations of their wave functions under continuous exchanges, leading to the possibility of braid anyons in 2D and traid anyons in 1D. Exclusion statistics on the other hand provides a completely different notion of anyons with fractional statistics. Originally proposed by Haldane [23, 35], later used by Wu [75] and Polychronakos [24, 76, 25] to formulate a general theory of quantum statistical mechanics, exclusion statistics defines anyons as particles obeying a generalized Pauli exclusion principle, interpolating between boson and fermions. Importantly, the theory makes no reference to the spatial dimension of the system, allowing for the existence of anyons not only in 2D (like it was historically assumed to be the case for exchange statistics), but also in 1D or 3D systems. Although the concepts of exclusion and exchange statistics in general do not give equivalent descriptions of fractional statistics, there are cases where both theories coincide, like for the quasiparticles in the fractional quantum Hall effect [23]. Various 1D systems exhibiting quasiparticles which obey a generalized Pauli principle are known in the literature. Famous examples include the Calogero-Sutherland model [77, 78], the Kundu model [42] and the Haldane-Shastry model [79]. Here we only give a brief review of the concept of exclusion statistics and refer to the literature for more details.

Consider Nisubscript𝑁𝑖N_{i}italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT identical particles of species i𝑖iitalic_i in a system of N𝑁Nitalic_N particles in total. When we fix the coordinates of N1𝑁1N-1italic_N - 1 of these particles, the single-particle Hilbert space dimension of the remaining particle of species i𝑖iitalic_i is denoted as di({Nj})subscript𝑑𝑖subscript𝑁𝑗d_{i}\left(\{N_{j}\}\right)italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( { italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT } ), which generally depends on the particle numbers of all species {Nj}subscript𝑁𝑗\{N_{j}\}{ italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT }. We assume that the single-particle Hilbert space dimension disubscript𝑑𝑖d_{i}italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is finite and extensive [23]. We now consider how disubscript𝑑𝑖d_{i}italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT changes when we add or subtract particles of different species (ΔNjΔsubscript𝑁𝑗\Delta N_{j}roman_Δ italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT) to the system. The change ΔdiΔsubscript𝑑𝑖\Delta d_{i}roman_Δ italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of the single-particle Hilbert space dimension is given by

Δdi=jαijΔNj,Δsubscript𝑑𝑖subscript𝑗subscript𝛼𝑖𝑗Δsubscript𝑁𝑗\displaystyle\Delta d_{i}=-\sum_{j}\alpha_{ij}\Delta N_{j},roman_Δ italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT roman_Δ italic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (57)

where Haldane defined the statistical interaction αijsubscript𝛼𝑖𝑗\alpha_{ij}italic_α start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT. In the case of bosons, the single-particle Hilbert space dimension does not change when particles of any species are added to the system and thus αij=0subscript𝛼𝑖𝑗0\alpha_{ij}=0italic_α start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = 0. Fermions on the other hand obey the Pauli principle. This means, if a fermion of species i𝑖iitalic_i is added to the system, the single-particle Hilbert space dimension disubscript𝑑𝑖d_{i}italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT gets reduced by 1, i.e. two fermions cannot occupy the same quantum state. In contrast, if a fermion of a different species j𝑗jitalic_j is added, disubscript𝑑𝑖d_{i}italic_d start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT doesn’t change. The statistical interaction of fermions is thus given by αij=δijsubscript𝛼𝑖𝑗subscript𝛿𝑖𝑗\alpha_{ij}=\delta_{ij}italic_α start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT. Quasiparticles where the statistical interaction is different from the boson or fermion case are called anyons with fractional statistics. The special case where αij0subscript𝛼𝑖𝑗0\alpha_{ij}\neq 0italic_α start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ≠ 0 for ij𝑖𝑗i\neq jitalic_i ≠ italic_j is also called mutual statistics. Building on Eq. 57, Wu derived the occupation numbers of a gas of ideal anyons at finite temperature without mutual statistics (αij=αδijsubscript𝛼𝑖𝑗𝛼subscript𝛿𝑖𝑗\alpha_{ij}=\alpha\delta_{ij}italic_α start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = italic_α italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT) [75]. For the mean occupation number nisubscript𝑛𝑖n_{i}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT at zero temperature Wu found

ni={0,if ϵi>EF1/α,if ϵi<EF,subscript𝑛𝑖cases0if ϵi>EF𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒1𝛼if ϵi<EF𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒\displaystyle n_{i}=\begin{cases}0,\quad\quad\text{if $\epsilon_{i}>E_{F}$}\\ 1/\alpha,\quad\text{if $\epsilon_{i}<E_{F}$}\end{cases},italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = { start_ROW start_CELL 0 , if italic_ϵ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT > italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL 1 / italic_α , if italic_ϵ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT < italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_CELL start_CELL end_CELL end_ROW , (58)

where ϵisubscriptitalic-ϵ𝑖\epsilon_{i}italic_ϵ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the single-particle energy of a particle in a state i𝑖iitalic_i and EFsubscript𝐸𝐹E_{F}italic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is the Fermi energy. Moreover, at finite temperature ni1/αsubscript𝑛𝑖1𝛼n_{i}\leq 1/\alphaitalic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≤ 1 / italic_α holds. We thus have a simple interpretation of the statistical interaction α𝛼\alphaitalic_α: For example in the case of α=0.5𝛼0.5\alpha=0.5italic_α = 0.5, every quantum state with energy ϵisubscriptitalic-ϵ𝑖\epsilon_{i}italic_ϵ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT can be occupied by at most 2 anyons. In the ground state all of those states up to the Fermi energy are occupied by exactly 2 anyons. The same holds for other values α=1/r𝛼1𝑟\alpha=1/ritalic_α = 1 / italic_r, where each state can hold up to r𝑟ritalic_r particles.

References

  • [1] J. M. Leinaas and J. Myrheim, On the theory of identical particles, Il Nuovo Cimento B 37(1), 1 (1977), 10.1007/BF02727953.
  • [2] G. A. Goldin, R. Menikoff and D. H. Sharp, Representations of a local current algebra in nonsimply connected space and the Aharonov–Bohm effect, J. Math. Phys. 22(8), 1664 (1981), 10.1063/1.525110.
  • [3] F. Wilczek, Quantum Mechanics of Fractional-Spin Particles, Phys. Rev. Lett. 49(14), 957 (1982), 10.1103/PhysRevLett.49.957.
  • [4] Y.-S. Wu, General Theory for Quantum Statistics in Two Dimensions, Phys. Rev. Lett. 52(24), 2103 (1984), 10.1103/PhysRevLett.52.2103.
  • [5] Y.-S. Wu, Multiparticle Quantum Mechanics Obeying Fractional Statistics, Phys. Rev. Lett. 53(2), 111 (1984), 10.1103/PhysRevLett.53.111.
  • [6] A. Khare, Fractional Statistics And Quantum Theory, World Scientific, Singapore, second edn., ISBN 978-981-256-775-8 (2005).
  • [7] D. Arovas, J. R. Schrieffer and F. Wilczek, Fractional Statistics and the Quantum Hall Effect, Phys. Rev. Lett. 53(7), 722 (1984), 10.1103/PhysRevLett.53.722.
  • [8] F. Wilczek, Fractional statistics and anyon superconductivity, World Scientific, Singapore, ISBN 978-981-02-0048-0 978-981-02-0049-7 (1990).
  • [9] H. Bartolomei, M. Kumar, R. Bisognin, A. Marguerite, J.-M. Berroir, E. Bocquillon, B. Plaçais, A. Cavanna, Q. Dong, U. Gennser, Y. Jin and G. Fève, Fractional statistics in anyon collisions, Science 368(6487), 173 (2020), 10.1126/science.aaz5601.
  • [10] J. Nakamura, S. Liang, G. C. Gardner and M. J. Manfra, Direct observation of anyonic braiding statistics, Nat. Phys. 16(9), 931 (2020), 10.1038/s41567-020-1019-1.
  • [11] T. Keilmann, S. Lanzmich, I. McCulloch and M. Roncaglia, Statistically induced phase transitions and anyons in 1D optical lattices, Nat. Comm. 2, 361 (2011), 10.1038/ncomms1353.
  • [12] S. Greschner and L. Santos, Anyon hubbard model in one-dimensional optical lattices, Phys. Rev. Lett. 115, 053002 (2015), 10.1103/PhysRevLett.115.053002.
  • [13] C. Sträter, S. C. L. Srivastava and A. Eckardt, Floquet realization and signatures of one-dimensional anyons in an optical lattice, Phys. Rev. Lett. 117, 205303 (2016), 10.1103/PhysRevLett.117.205303.
  • [14] J. Kwan, P. Segura, Y. Li, S. Kim, A. V. Gorshkov, A. Eckardt, B. Bakkali-Hassani and M. Greiner, Realization of 1d anyons with arbitrary statistical phase (2023), 2306.01737.
  • [15] M. Bonkhoff, K. Jägering, S. Eggert, A. Pelster, M. Thorwart and T. Posske, Bosonic continuum theory of one-dimensional lattice anyons, Phys. Rev. Lett. 126, 163201 (2021), 10.1103/PhysRevLett.126.163201.
  • [16] N. L. Harshman and A. C. Knapp, Anyons from three-body hard-core interactions in one dimension, Ann. Phys. 412, 168003 (2020), 10.1016/j.aop.2019.168003.
  • [17] N. L. Harshman and A. C. Knapp, Topological exchange statistics in one dimension, Phys. Rev. A 105, 052214 (2022), 10.1103/PhysRevA.105.052214.
  • [18] M. Khovanov, Doodle Groups, Trans. Am. Math. Soc. 349(6), 2297 (1997).
  • [19] A. Bartholomew, R. Fenn, N. Kamada and S. Kamada, Doodles on surfaces, J. Knot Theory Ramifications 27(12) (2018).
  • [20] V. Bardakov, M. Singh and A. Vesnin, Structural aspects of twin and pure twin groups, Geometriae Dedicata 203, 135 (2019).
  • [21] T. Naik, N. Nanda and M. Singh, Conjugacy classes and automorphisms of twin groups, Forum Mathimaticum 32(5), 1095 (2020).
  • [22] J. González, J. L. León-Medina and C. Roque-Márquez, Linear motion planning with controlled collisions and pure planar braids, Homology, Homotopy and Applications 23(1), 275 (2021), 10.4310/HHA.2021.v23.n1.a15.
  • [23] F. D. M. Haldane, “Fractional statistics” in arbitrary dimensions: A generalization of the Pauli principle, Phys. Rev. Lett. 67(8), 937 (1991), 10.1103/PhysRevLett.67.937.
  • [24] A. P. Polychronakos, Probabilities and path-integral realization of exclusion statistics, Physics Letters B 365(1), 202 (1996), https://doi.org/10.1016/0370-2693(95)01302-4.
  • [25] A. P. Polychronakos, Generalized statistics in one dimension, In A. Comtet, T. Jolicœur, S. Ouvry and F. David, eds., Aspects topologiques de la physique en basse dimension. Topological aspects of low dimensional systems, pp. 415–471. Springer Berlin Heidelberg, Berlin, Heidelberg, ISBN 978-3-540-46637-6 (1999).
  • [26] N. L. Harshman, One-Dimensional Traps, Two-Body Interactions, Few-Body Symmetries. II. N𝑁Nitalic_N particles, Few-Body Systems 57(1), 45 (2016), 10.1007/s00601-015-1025-5.
  • [27] N. L. Harshman, One-Dimensional Traps, Two-Body Interactions, Few-Body Symmetries. I. 1,2, and 3 particles, Few-Body Systems 57(1), 11 (2016), 10.1007/10.1007/s00601-015-1024-6.
  • [28] F. Lange, S. Ejima and H. Fehske, Strongly repulsive anyons in one dimension, Phys. Rev. A 95, 063621 (2017), 10.1103/PhysRevA.95.063621.
  • [29] J. Friedel, Metallic alloys, Il Nuovo Cimento 7(S2), 287 (1958), 10.1007/bf02751483.
  • [30] S. Greschner, L. Cardarelli and L. Santos, Probing the exchange statistics of one-dimensional anyon models, Phys. Rev. A 97, 053605 (2018), 10.1103/PhysRevA.97.053605.
  • [31] F. H. L. Essler, H. Frahm, F. Göhmann, A. Klümper and V. E. Korepin, The One-Dimensional Hubbard Model, Cambridge University Press, ISBN 978-1-139-44158-2 (2005).
  • [32] P. Blasiak, A. Horzela, K. A. Penson, A. I. Solomon and G. H. E. Duchamp, Combinatorics and boson normal ordering: A gentle introduction, Am. J. Phys. 75(7), 639 (2007), 10.1119/1.2723799.
  • [33] A. K. Kolezhuk, F. Heidrich-Meisner, S. Greschner and T. Vekua, Frustrated spin chains in strong magnetic field: Dilute two-component bose gas regime, Phys. Rev. B 85, 064420 (2012), 10.1103/PhysRevB.85.064420.
  • [34] M. Greiter, Statistical phases and momentum spacings for one-dimensional anyons, Phys. Rev. B 79, 064409 (2009), 10.1103/PhysRevB.79.064409.
  • [35] M. Greiter and F. Wilczek, Fractional statistics, 10.48550/ARXIV.2210.02530 (2022).
  • [36] B. Paredes, T. Keilmann and J. I. Cirac, Pfaffian-like ground state for three-body hard-core bosons in one-dimensional lattices, Phys. Rev. A 75(5), 053611 (2007), 10.1103/PhysRevA.75.053611.
  • [37] L. Mazza, M. Rizzi, M. Lewenstein and J. I. Cirac, Emerging bosons with three-body interactions from spin-1 atoms in optical lattices, Phys. Rev. A 82, 043629 (2010), 10.1103/PhysRevA.82.043629.
  • [38] N. L. Harshman, Symmetries of three harmonically trapped particles in one dimension, Phys. Rev. A 86, 052122 (2012), 10.1103/PhysRevA.86.052122.
  • [39] M. Girardeau, Relationship between Systems of Impenetrable Bosons and Fermions in One Dimension, J. Math. Phys. 1(6), 516 (1960), doi:10.1063/1.1703687.
  • [40] T. Cheon and T. Shigehara, Fermion-boson duality of one-dimensional quantum particles with generalized contact interactions, Phys. Rev. Lett. 82, 2536 (1999), 10.1103/PhysRevLett.82.2536.
  • [41] M. Valiente, Bose-fermi dualities for arbitrary one-dimensional quantum systems in the universal low-energy regime, Phys. Rev. A 102, 053304 (2020), 10.1103/PhysRevA.102.053304.
  • [42] A. Kundu, Exact Solution of Double $\ensuremath{\delta}$ Function Bose Gas through an Interacting Anyon Gas, Phys. Rev. Lett. 83(7), 1275 (1999), 10.1103/PhysRevLett.83.1275.
  • [43] M. T. Batchelor, X.-W. Guan and N. Oelkers, One-Dimensional Interacting Anyon Gas: Low-Energy Properties and Haldane Exclusion Statistics, Phys. Rev. Lett. 96(21), 210402 (2006), 10.1103/PhysRevLett.96.210402.
  • [44] O. I. Pâţu, V. E. Korepin and D. V. Averin, Correlation functions of one-dimensional lieb–liniger anyons, J. Phys. A 40(50), 14963 (2007), 10.1088/1751-8113/40/50/004.
  • [45] T. Posske, B. Trauzettel and M. Thorwart, Second quantization of leinaas-myrheim anyons in one dimension and their relation to the lieb-liniger model, Phys. Rev. B 96, 195422 (2017), 10.1103/PhysRevB.96.195422.
  • [46] O. I. Pâţu, Correlation functions of one-dimensional strongly interacting two-component gases, Phys. Rev. A 100, 063635 (2019), 10.1103/PhysRevA.100.063635.
  • [47] L. Piroli, S. Scopa and P. Calabrese, Determinant formula for the field form factor in the anyonic Lieb–Liniger model, J. Phys. A 53(40), 405001 (2020), 10.1088/1751-8121/ab94ed.
  • [48] S. J. Benetton Rabello, 1d generalized statistics gas: A gauge theory approach, Phys. Rev. Lett. 76, 4007 (1996), 10.1103/PhysRevLett.76.4007.
  • [49] U. Aglietti, L. Griguolo, R. Jackiw, S.-Y. Pi and D. Seminara, Anyons and chiral solitons on a line, Phys. Rev. Lett. 77, 4406 (1996), 10.1103/PhysRevLett.77.4406.
  • [50] L. Griguolo and D. Seminara, Chiral solitons from dimensional reduction of chern-simons gauged non-linear schrödinger equation: classical and quantum aspects, Nuclear Physics B 516(1), 467 (1998), https://doi.org/10.1016/S0550-3213(97)00810-9.
  • [51] C. S. Chisholm, A. Frölian, E. Neri, R. Ramos, L. Tarruell and A. Celi, Encoding a one-dimensional topological gauge theory in a raman-coupled bose-einstein condensate, Phys. Rev. Res. 4, 043088 (2022), 10.1103/PhysRevResearch.4.043088.
  • [52] A. Frölian, C. S. Chisholm, E. Neri, C. R. Cabrera, R. Ramos, A. Celi and L. Tarruell, Realizing a 1D topological gauge theory in an optically dressed BEC, Nature 608(7922), 293 (2022), 10.1038/s41586-022-04943-3.
  • [53] J. S. Dowker, Quantum mechanics and field theory on multiply connected and on homogeneous spaces, J. Phys. A 5(7), 936 (1972), 10.1088/0305-4470/5/7/004.
  • [54] C. J. Isham, TOPOLOGICAL AND GLOBAL ASPECTS OF QUANTUM THEORY, In Les Houches Summer School on Theoretical Physics: Relativity, Groups and Topology, pp. 1059–1290 (1983).
  • [55] T. D. Imbo and E. C. G. Sudarshan, Inequivalent quantizations and fundamentally perfect spaces, Phys. Rev. Lett. 60(6), 481 (1988), 10.1103/PhysRevLett.60.481.
  • [56] T. D. Imbo, C. Shah Imbo and E. C. G. Sudarshan, Identical particles, exotic statistics and braid groups, Physics Letters B 234(1), 103 (1990), 10.1016/0370-2693(90)92010-G.
  • [57] A. Balachandran, Classical topology and quantum statistics, Int. J. Mod. Phys. B 05(16n17), 2585 (1991), 10.1142/S0217979291001036.
  • [58] L. Schulman, A Path Integral for Spin, Phys. Rev. 176(5), 1558 (1968), 10.1103/PhysRev.176.1558.
  • [59] M. G. G. Laidlaw and C. M. DeWitt, Feynman Functional Integrals for Systems of Indistinguishable Particles, Phys. Rev. D 3(6), 1375 (1971), 10.1103/PhysRevD.3.1375.
  • [60] M. Bourdeau and R. D. Sorkin, When can identical particles collide?, Phys. Rev. D 45(2), 687 (1992), 10.1103/PhysRevD.45.687.
  • [61] N. P. K. Landsman, Quantization and superselection sectors III: Multiply connected spaces and indistinguishable particles, Reviews in Mathematical Physics 28(09), 1650019 (2016), 10.1142/S0129055X16500197.
  • [62] S. Ohya, A generalization of the one-dimensional boson–fermion duality through the path-integral formalism, Annals of Physics 434, 168657 (2021), https://doi.org/10.1016/j.aop.2021.168657.
  • [63] S. Ohya, Discrete scale-invariant boson-fermion duality in one dimension, Phys. Rev. A 105, 033312 (2022), 10.1103/PhysRevA.105.033312.
  • [64] X.-G. Wen, Quantum Field Theory of Many-body Systems: From the Origin of Sound to an Origin of Light and Electrons, Oxford Graduate Texts. Oxford University Press, Oxford, New York, ISBN 978-0-19-922725-9 (2007).
  • [65] W. Thurston, Geometry and topology of three-manifolds, unpublished, electronic version 1.1 edn. (2002).
  • [66] A. Adem, J. Leida and Y. Ruan, Orbifolds and Stringy Topology, Cambridge University Press, Cambridge ; New York, 1st edition edn., ISBN 978-0-521-87004-7 (2007).
  • [67] F. Wilczek, Magnetic Flux, Angular Momentum, and Statistics, Phys. Rev. Lett. 48(17), 1144 (1982), 10.1103/PhysRevLett.48.1144.
  • [68] Z. N. C. Ha, Fractional statistics in one dimension: view from an exactly solvable model, Nuc. Phys. B 435(3), 604 (1995), 10.1016/0550-3213(94)00537-O.
  • [69] S. Sree Ranjani, P. Panigrahi, A. Kapoor and A. Khare, An explicit realization of fractional statistics in one dimension, Ann. Phys. 324(5), 1176 (2009), https://doi.org/10.1016/j.aop.2009.01.007.
  • [70] M. D. Girardeau, Anyon-Fermion Mapping and Applications to Ultracold Gases in Tight Waveguides, Phys. Rev. Lett. 97(10), 100402 (2006), 10.1103/PhysRevLett.97.100402.
  • [71] A. del Campo, Fermionization and bosonization of expanding one-dimensional anyonic fluids, Phys. Rev. A 78, 045602 (2008), 10.1103/PhysRevA.78.045602.
  • [72] N. E. Mackel, J. Yang and A. del Campo, Quantum alchemy and universal orthogonality catastrophe in one-dimensional anyons, 10.48550/ARXIV.2210.10776 (2022), 2210.10776.
  • [73] G. Moore and N. Read, Nonabelions in the fractional quantum hall effect, Nuc. Phys. B 360(2), 362 (1991), 10.1016/0550-3213(91)90407-O.
  • [74] C. Nayak, S. H. Simon, A. Stern, M. Freedman and S. Das Sarma, Non-Abelian anyons and topological quantum computation, Rev. Mod. Phys. 80(3), 1083 (2008), 10.1103/RevModPhys.80.1083.
  • [75] Y.-S. Wu, Statistical distribution for generalized ideal gas of fractional-statistics particles, Physical Review Letters 73(7), 922 (1994), 10.1103/physrevlett.73.922.
  • [76] S. B. Isakov, D. P. Arovas, J. Myrheim and A. P. Polychronakos, Thermodynamics for fractional exclusion statistics, 10.48550/ARXIV.COND-MAT/9601108 (1996), cond-mat/9601108.
  • [77] F. Calogero, Ground state of a one-dimensional n body system, Journal of Mathematical Physics 10(12), 2197 (1969), 10.1063/1.1664821.
  • [78] B. Sutherland, Quantum many-body problem in one dimension: Ground state, Journal of Mathematical Physics 12(2), 246 (1971), 10.1063/1.1665584.
  • [79] S. Chaturvedi and V. Srinivasan, Microscopic interpretation of haldane’s semion statistics, Phys. Rev. Lett. 78, 4316 (1997), 10.1103/PhysRevLett.78.4316.