Abstract
Quantum statistics plays a fundamental role in the laws of nature. Haldane fractional exclusion statistics (FES) generalizes the Pauli exclusion statistics, and can emerge in the properties of elementary particles and hole excitations of a quantum system consisting of conventional bosons or fermions. FES has a long history of intensive studies, but its simple realization in interacting physical systems is rare. Here we report a simple non-mutual FES that depicts the particle-hole symmetry breaking in interacting Bose gases at a quantum critical point. We show that the FES distribution directly comes from particle-hole symmetry breaking. Based on exact solutions, quantum Monte Carlo simulations and experiments, we find that, over a wide range of interaction strengths, the macroscopic physical properties of these gases are determined by non-interacting quasi-particles that obey non-mutual FES of the same form in one and two dimensions. Whereas strongly interacting Bose gases reach full fermionization in one dimension, they exhibit incomplete fermionization in two dimensions. Our results provide a generic connection between interaction-induced particle-hole symmetry breaking (depicted by FES) and macroscopic properties of many-body systems in arbitrary dimensions. Our work lays the groundwork for using FES to explore quantum criticality and other novel many-body phenomena in strongly correlated quantum systems.
Keywords: quantum statistics, interaction, particle-hole symmetry breaking, fractional exclusion statistics, strongly correlated quantum materials
Near a quantum critical point, the macroscopic properties of interacting many-body Bose gases in one and two dimensions are determined by ideal particles that obey non-mutual fractional exclusion statistics.
INTRODUCTION
Bose-Einstein and Fermi-Dirac statistics constitute two cornerstones of quantum statistical mechanics. However, they are not the only possible forms of quantum statistics [1]. In two dimensions, anyonic excitations can carry fractional charges and obey fractional statistics [2–7]. To generalize fractional statistics, Haldane formulated a theory of fractional exclusion statistics (FES) that continuously interpolates between Bose and Fermi statistics in arbitrary spatial dimensions [8]. This theory depicts how much the Hilbert space dimensionality for available single-particle states, namely the ‘number of holes’ (Nh, α) of species α decreases as particles of species β are added to a system [8–10]:
![]() |
(1) |
Here the FES parameter gαβ is independent of the particle number NP, β. Bose and Fermi statistics correspond to the non-mutual FES where gαβ = gδαβ with g = 0 and 1, respectively. FES has found exact realizations in a few one-dimensional (1D) systems, including the Calogero-Sutherland model of particles interacting through a 1/r2 potential [11–14], Lieb-Liniger Bose gases [14,15] and anyonic gases with delta-function interaction [16,17].
FES reveals the statistical nature of a system with respect to its energy spectrum regardless of whether the constituent particles interact or not. On the other hand, particle-hole symmetry breaking (PHSB) [18] emerges as a key mechanism for understanding strongly correlated quantum materials including high-Tc superconductors [19,20] and fractional quantum Hall systems [21]. This symmetry breaking significantly influences physical properties such as equations of state [22], optical properties [23], dynamical evolutions [24], transport properties [25] and non-Fermi-liquid behaviors [26]. However, it remains challenging to identify emergent FES for depicting the particle-hole symmetry breaking in generic interacting many-body systems.
In this article, we show that FES naturally emerges as a result of particle-hole symmetry breaking in quantum many-body systems (see Fig. 1(a) and Equation (2)). In particular, we demonstrate interaction-induced non-mutual FES at a quantum critical point. We consider a repulsively interacting Bose gas that undergoes a quantum phase transition under zero temperature (T = 0) from a vacuum to a quantum liquid when the chemical potential μ exceeds a critical value μc (Fig. 1(b)). Here ‘quantum liquid’ denotes a Tomonaga-Luttinger liquid (TLL) [27] in one dimension or a superfluid in higher dimensions [28]. Based on exact solutions in one dimension and high-precision quantum Monte Carlo (QMC) simulations in two dimensions, we report evidence for emergent particle-hole symmetry breaking (Equation (2)) in these many-body systems. Our results are further supported by existing experimental data given in [27–32]. We establish a one-to-one correspondence between interaction and particle-hole symmetry breaking over a wide range of interaction strengths, and further observe that, remarkably, such symmetry breaking determines the macroscopic properties of interacting gases in a unified manner, as summarized by the logic flow shown in Fig. 1(c).
Figure 1.

Particle-hole symmetry breaking and Haldane FES in interacting many-body systems. (a) Particle-hole symmetry breaking in a quasi-momentum cell at
(see Equation (2)) in an interacting system. (b) A many-body system near a quantum critical point. (c) Schematic of the logic flow: interaction determines the particle-hole symmetry breaking (depicted by FES) that in turn governs the macroscopic physical properties of a many-body system at and near a quantum critical point.
Particle-hole symmetry breaking and FES
In a quantum many-body system, interaction dresses the constituent particles to form quasi-particles that are statistically distributed over the quasi-momentum space. In each quasi-momentum cell that defines the species in Equation (1), the number of unoccupied states depends on the numbers of occupied states in the same cell and in other cells, forming a net of correlated cells in general. Such correlated cells can be depicted by quasi-momentum-dependent particle-hole symmetry breaking; see Equation S43 within the online supplementary material. At a quantum critical point where the system is strongly correlated with large characteristic lengths in real space, these quasi-momentum cells become decoupled into nearly independent cells (see Section 3 within the online supplementary material), which leads to a simplified form of particle-hole symmetry breaking equation that we expect to hold in arbitrary dimensions (Fig. 1(a)):
![]() |
(2) |
Here, the species label α is given by quasi-momentum
, gαβ by
, NP and Nh are scaled into distribution functions,
and
, of occupied states and of holes, and dsp = 1/(2π)D is a bare dimensionality of states in a phase-space unit cell for a D-dimensional system.
Accordingly, a non-mutual FES distribution [9,14] of quasi-particles naturally emerges at a quantum critical point. We prove (for details, see Section 4 within the online supplementary material) that Equation (2) directly gives rise to a non-mutual FES distribution with the following occupation number f in a state with energy ε:
![]() |
(3) |
This generic connection between Equations (2) and (3) enables understanding interacting systems from the perspective of emergent FES phenomena. Based on the interaction-induced particle-hole symmetry breaking (Equation (2)), the macroscopic physical properties of many-body systems can be obtained through non-interacting quasi-particles that obey the non-mutual FES distribution in Equation (3). Specifically, the number density and energy density are given by n = ∫G(ε)f(ε)dε and e = ∫G(ε)f(ε)εdε, where the density of states per volume is given by
in one dimension and 1/(4π) in two dimensions for non-relativistic particles. We set 2m = kB = ℏ = 1, where m is the particle mass, kB the Boltzmann constant and ℏ the reduced Planck constant.
To connect Equation (2) to physical interacting systems, we introduce an interaction-FES correspondence hypothesis:
![]() |
(4) |
Here
is a properly scaled interaction strength,
is the corresponding FES parameter in Equations (2) and (3),
and
is a coefficient. This hypothesis is inspired by an analytical result for the FES in 1D strongly interacting Bose gases [33] as well as a Ginzburg-Landau theory for 2D superfluids [31,34], and is found to apply over a large interaction range (see Fig. 2). It provides a simple proportionality relation between the interaction strength
and the resulting Hilbert space dimensionality ratio (
) of g − g0 (for the Hilbert space occupied by one single particle because of interaction) to g∞ − g (for the ‘remaining’ Hilbert space that is occupiable but yet unoccupied because
has not reached infinity). For interacting Bose gases, g0 = 0 and we denote g∞ as gmax. Equations (2) and (4) together enable quantitative predictions of macroscopic physical properties. In the following, we provide evidence for the emergence of such interaction-induced particle-hole symmetry breaking and non-mutual FES at and near a quantum critical point.
Figure 2.

Demonstration of the proportionality relation in the interaction-FES correspondence hypothesis (Equation (4)). Based on exact solutions for 1D Bose gases with delta-function interaction, we compute
that depicts the particle-hole symmetry breaking in low-energy excitations. The numerical data (circles) show the proportionality relation with fitted
. The solid line shows Equation (4) with c1 = 0.772 (see Equation (10) for details). Under strong couplings (
), the numerical data can be described by an analytical form (dashed line; see Equation (7)).
FES in one dimension
The 1D Bose gas with delta-function interaction is an integrable model with great relevance in both theoretical and experimental contexts (see reviews [35,36]). Such gases are described by the Hamiltonian [15,37]
![]() |
(5) |
where c is the repulsive elastic interaction strength and N the particle number. In its dilution limit, the discrete 1D Bose-Hubbard model used in QMC simulations corresponds to Equation (5) with c = U/(2t1/2) (see Section 9 within the online supplementary material), where U and t are the onsite interaction and tunneling parameters, respectively. We exactly solve such 1D gases at the vacuum-to-TLL transition (μc = 0) [27] based on the thermodynamic Bethe ansatz (TBA) equation [37,38],
![]() |
(6) |
where a(x) = 2c/(c2 + x2), and the pressure is given by p(μ, T) = (T/2π)∫ln (1 + e−ϵ(k)/T)dk. For convenience, we present thermodynamic observables and parameters in dimensionless forms (see Section 1 within the online supplementary material). We compute the critical entropy per particle Sc/N ≡ (S/N)(μ = μc), scaled critical density
and scaled critical pressure
by numerically solving Equation (6).
The Sc/N increases with a scaled interaction strength
(Fig. 3(a)). It reaches A∞, 1D ≈ 1.89738 at
(see Section 1 within the online supplementary material), exactly matching the Sc/N of non-interacting fermions [37] (gmax, 1D = 1), as predicted and observed for Tonks-Girardeau gases [15,39–42]. These solutions agree with data extracted from experiments performed by the Kaiserslautern group [29] and the USTC group [27], and agree with our 1D QMC simulations (see Sections 10 and 11 within the online supplementary material).
Figure 3.

Evidence for interaction-induced FES in 1D Bose gases at a quantum critical point. (a) Critical entropy per particle Sc/N as a function of
. Exact solutions (circles) agree excellently with QMC computations (diamonds) and agree with experiments (squares and triangles, from [27,29]). (b) Power-law scaling of Sc/N with respect to
. Dotted line denotes the fermionization limit A∞, 1D. (c)–(e) Under Equation (8) with gmax, 1D = 1, thermodynamic observables of interacting gases agree well with those of non-interacting quasi-particles that obey non-mutual FES: (c) Sc/N; (d) scaled critical density
; (e) scaled critical pressure
. Error bars represent 1σ statistical uncertainties.
We now verify the particle-hole symmetry breaking equation (Equation (2)) and the interaction-FES correspondence hypothesis (Equation (4)) based on ab initio computations (see Section 5 within the online supplementary material). We compute
and
and define a
-dependent FES parameter
based on
. We observe that
is almost homogeneous within a finite range of
(see Section 5 within the online supplementary material), corresponding to low-energy elementary excitations that obey simple particle-hole symmetry breaking (Equation (2)) depicted by a non-mutual FES distribution (Equation (3)). Equation (3) associated with
captures the essential behaviors of
and
(see Section 5 within the online supplementary material). Furthermore, the Hilbert space dimensionality ratio
shows a proportionality relation to the interaction strength
over a large range (Fig. 2), with a fitted coefficient
. Thus, Equation (4) provides a powerful approximation that depicts the particle-hole symmetry breaking for low-energy excitations in interacting gases. We note that, based on Equation (5), particle-hole symmetry breaking can in general be depicted by
-dependent mutual FES [14]; such mutual FES reduces to non-mutual FES under strong coupling [33], which we here derive to be for finite temperatures (see Section 5 within the online supplementary material)
![]() |
(7) |
As shown in Fig. 2, Equation (4) not only agrees with this strong-coupling analytical form, but also depicts well the numerically computed
over a significantly larger range covering strong, intermediate and weak interactions.
The particle-hole symmetry breaking (depicted by Equations (2) and (4)) dictates the distribution functions (Equation (3)) of elementary excitations and thereby determines the macroscopic properties of interacting gases. For Bose gases, Equation (4) predicts a one-to-one mapping to non-interacting quasi-particles with FES parameter
![]() |
(8) |
where
is a transformed interaction parameter,
![]() |
(9) |
and gmax, 1D = 1. Based on the computed critical entropy per particle, we observe two scaling functions (Fig. 3(b) and (c)) that are similar to each other, which is characteristic of the interaction-FES correspondence (Equation (8)). For non-interacting FES quasi-particles, the Sc, FES/N at μc = 0 exhibits a power-law scaling with respect to g (Fig. 3(c), blue curve):
, with βFES, 1D = 0.298(2) fitted for 0.05 < g ≤ 1. This scaling suggests, and we indeed observe, that the Sc/N for interacting Bose gases accordingly obeys a power-law scaling with respect to
(Fig. 3(b)):
![]() |
(10) |
Here β1D = 0.298(1) and
are fitted parameters. Under Equation (8), these two scaling functions match each other, and the numerical data agree within 4% (Fig. 3(c)). We note that Equation (10) agrees excellently with exact solutions within
over
, and the precisely determined
conforms well to the
determined earlier. Thus, identifying the power-law scaling for Sc/N provides a smoking-gun signature and a precise determination of the interaction-FES correspondence.
We find agreement within
and
for
and
, respectively (Fig. 3(d) and (e)). The overall good agreement for Sc/N,
and
shows that macroscopic thermodynamic observables are determined by interaction-induced particle-hole symmetry breaking. Therefore, these observables can serve as a practical gauge for conveniently measuring the corresponding non-mutual FES, especially when ab initio computations are difficult or unavailable.
FES in two dimensions
In higher dimensions, while the Bethe ansatz in general does not apply, particle-hole symmetry breaking remains a key characteristic [43] that governs measurable macroscopic properties of interacting gases. Here we benchmark the applicability of Equations (2) and (4) in 2D gases. Using QMC simulations [44,45], we study a 2D Bose-Hubbard lattice gas that has a vacuum-to-superfluid quantum phase transition at μc = −4t [28]. The Bose-Hubbard Hamiltonian is given by
![]() |
(11) |
where
and
are the creation and annihilation operators at site i,
and 〈i, j〉 runs over all nearest neighboring sites. We define a scaled interaction strength
(see Section 9 within the online supplementary material) that is the lattice-gas equivalence [28,31] of the interaction parameter
for 2D Bose gases without lattices, where a is the scattering length and lz is an oscillator length [32].
To obtain physical properties that are insensitive to the lattice structure, we perform QMC simulations for each
at a series of temperatures down to T = 0.1t. We extract scaled quantities Sc/N,
and
for each T, and perform extrapolation towards T = 0 for each quantity (see Section 10 within the online supplementary material). We test this extrapolation protocol on a 1D Bose-Hubbard system (see Section 10 within the online supplementary material) and find excellent agreement with exact solutions (Fig. 3).
In two dimensions, we identify the same interaction-FES correspondence (Equation (8)) as in one dimension. The Sc/N increases with
and reaches A∞, 2D = 1.988(14) at
(Fig. 4(a)), matching the Sc, FES/N of non-interacting FES quasi-particles with gmax, 2D = 0.432(14). Our QMC data agree well with a non-perturbative renormalization group (NPRG) computation [46], and with experiments by the Chicago [28] and ENS [30] groups. Based on Equation (9) and
, Sc/N shows an excellent power-law scaling with respect to
(Fig. 4(b)):
![]() |
(12) |
with A∞, 2D,
, and β2D = 0.20(1) fitted for
(
). We then test the predictions of Equations (2) and (4) (with gmax, 2D and
determined above) using the QMC data.
Figure 4.

Evidence for interaction-induced FES in 2D Bose gases at a quantum critical point. (a) Critical entropy per particle Sc/N as a function of
. QMC results (circles) agree with NPRG computations [46] and experiments [28,30]. (b) Power-law scaling of Sc/N with respect to
. (c)–(e) Given Equation (8) with gmax, 2D = 0.432(14), thermodynamic observables of interacting gases agree well with those of non-interacting quasi-particles that obey non-mutual FES: (c) Sc/N; (d) scaled critical density
; (e) scaled critical pressure
, with the dotted line denoting the non-interacting boson limit
[46]. Our results agree with existing experiments [28,30,31]. Horizontal and vertical gray bands mark A∞, 2D and gmax, 2D, respectively. Error bars represent 1σ statistical uncertainties.
We find evidence for emergent non-mutual FES based on good agreement for Sc/N,
and
. The FES quasi-particles also exhibit a power-law scaling,
, with AFES, 2D ≡ (Sc, FES/N)(g = 1) ≈ 2.373. The exponent βFES, 2D = 0.2122(1) is fitted for 0.02 ≤ g ≤ 1 and agrees with β2D. Given Equation (8), the two power-law scaling functions for Sc/N versus
and for Sc, FES/N versus g agree well within
(Fig. 4(c)). Accordingly,
and
show agreement within
and
, respectively (Fig. 4(d) and (e)). Our simulations agree with existing experiments [28,30,31] at and near the quantum critical point (see also Fig. S5 within the online supplementary material). These numerical and experimental data together are fully consistent with, and thereby strongly support, the emergence of interaction-induced particle-hole symmetry breaking and non-mutual FES in two dimensions (Equations (2) and (8)). The less-than-unity gmax, 2D = 0.432(14) shows incomplete fermionization of strongly interacting 2D Bose gases.
In summary, we established a generic connection between particle-hole symmetry breaking and FES and then found strong evidence for interaction-induced non-mutual FES at and near a quantum critical point. Our non-perturbative approach holds promise for studying the dynamical evolutions [24], transport properties [25] and hydrodynamics [47] of quantum systems in arbitrary dimensions, and for studying other integrable models such as multi-component systems. In particular, such simple non-mutual FES is expected to emerge in the charge degree of freedom at the quantum critical region of the multi-component ultracold atoms in one and higher dimensions, whereas the spin degrees of freedom are frozen out at quantum criticality. Moreover, our approach provides a route to understanding strongly interacting quantum materials where experiments can be both enriched and complicated by inelastic collisional losses and finite temperature effects [31,48,49].
METHODS
Quantum Monte Carlo simulations
In our work, we apply the worm algorithm in the path-integral representation to simulate the Bose-Hubbard model in both one and two dimensions using the QMC method. Here, we mainly focus on three observables: particle density n, pressure p and entropy per particle S/N. Our QMC data are obtained by spending about 2 × 105 CPU hours. Further details of the QMC simulation can be found in Sections 8 to 11 within the online supplementary material.
Statistical methods for data analysis
The error bars in figures and texts represent 1σ statistical uncertainties of the QMC simulations or experimental measurements. Based on a set of numerical or experimental data and a specific model, a least-square fit can be performed to determine the best fitting parameters as well as the standard errors of these parameters.
DATA AVAILABILITY
The data supporting the findings of this study are available within the paper and accompanying online supplementary material.
Supplementary Material
Contributor Information
Xibo Zhang, International Center for Quantum Materials, School of Physics, Peking University, Beijing 100871, China; Collaborative Innovation Center of Quantum Matter, Beijing 100871, China; Beijing Academy of Quantum Information Sciences, Beijing 100193, China.
Yang-Yang Chen, State Key Laboratory of Magnetic Resonance and Atomic and Molecular Physics, Wuhan Institute of Physics and Mathematics, Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences, Wuhan 430071, China; Institute of Modern Physics, Northwest University, Xi’an 710127, China.
Longxiang Liu, Chinese Academy of Sciences Center for Excellence in Quantum Information and Quantum Physics, University of Science and Technology of China, Hefei 230326, China.
Youjin Deng, Chinese Academy of Sciences Center for Excellence in Quantum Information and Quantum Physics, University of Science and Technology of China, Hefei 230326, China; MinJiang Collaborative Center for Theoretical Physics, College of Physics and Electronic Information Engineering, Minjiang University, Fuzhou 350108, China.
Xiwen Guan, State Key Laboratory of Magnetic Resonance and Atomic and Molecular Physics, Wuhan Institute of Physics and Mathematics, Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences, Wuhan 430071, China; Department of Theoretical Physics, Research School of Physics and Engineering, Australian National University, Canberra ACT 0200, Australia.
ACKNOWLEDGEMENTS
We are grateful to Cheng Chin for insightful discussions. We acknowledge Chun-Jiong Huang, Zhen-Sheng Yuan, Chen-Lung Hung, Li-Chung Ha, Biao Wu, Hui Zhai, Shina Tan for discussions and technical support.
FUNDING
This work was supported by the National Key Research and Development Program of China (2018YFA0305601, 2016YFA0300901, 2017YFA0304500 and 2016YFA0301604), the National Natural Science Foundation of China (11874073, 12134015, 11874393, 11625522 and 12104372), the Chinese Academy of Sciences Strategic Priority Research Program (XDB35020100), the Chinese Academy of Sciences Innovation Team (12121004), and the Hefei National Laboratory.
AUTHOR CONTRIBUTIONS
X.Z., Y.D. and X.G. conceived the project and wrote the paper. Y.-Y.C. and L.L. performed the numerical and analytical studies on the 1D and 2D models, respectively. X.Z. also performed analytical and numerical computations for this paper.
Conflict of interest statement. None declared.
REFERENCES
- 1. Khare A. Fractional Statistics and Quantum Theory. Singapore: World Scientific Publishing, 2005. 10.1142/5752 [DOI] [Google Scholar]
- 2. Leinaas JM, Myrheim J. On the theory of identical particles. Nuovo Cim B 1977; 37: 1–23. 10.1007/BF02727953 [DOI] [Google Scholar]
- 3. Wilczek F. Magnetic flux, angular momentum, and statistics. Phys Rev Lett 1982; 48: 1144–6. 10.1103/PhysRevLett.48.1144 [DOI] [Google Scholar]
- 4. Wilczek F. Quantum mechanics of fractional-spin particles. Phys Rev Lett 1982; 49: 957–9. 10.1103/PhysRevLett.49.957 [DOI] [Google Scholar]
- 5. Arovas D, Schrieffer JR, Wilczek F. Fractional statistics and the quantum Hall effect. Phys Rev Lett 1984; 53: 722–3. 10.1103/PhysRevLett.53.722 [DOI] [Google Scholar]
- 6. Laughlin RB. Superconducting ground state of noninteracting particles obeying fractional statistics. Phys Rev Lett 1988; 60: 2677–80. 10.1103/PhysRevLett.60.2677 [DOI] [PubMed] [Google Scholar]
- 7. Laughlin RB. The relationship between high-temperature superconductivity and the fractional quantum Hall effect. Science 1988; 242: 525–33. 10.1126/science.242.4878.525 [DOI] [PubMed] [Google Scholar]
- 8. Haldane FDM. Fractional statistics in arbitrary dimensions: a generalization of the Pauli principle. Phys Rev Lett 1991; 67: 937–40. 10.1103/PhysRevLett.67.937 [DOI] [PubMed] [Google Scholar]
- 9. Wu YS. Statistical distribution for generalized ideal gas of fractional-statistics particles. Phys Rev Lett 1994; 73: 922–5. 10.1103/PhysRevLett.73.922 [DOI] [PubMed] [Google Scholar]
- 10. Isakov SB. Statistical mechanics for a class of quantum statistics. Phys Rev Lett 1994; 73: 2150–3. 10.1103/PhysRevLett.73.2150 [DOI] [PubMed] [Google Scholar]
- 11. Calogero F. Solution of a three-body problem in one dimension. J Math Phys 1969; 10: 2191–6. 10.1063/1.1664820 [DOI] [Google Scholar]
- 12. Calogero F. Ground state of a one-dimensional n-body system. J Math Phys 1969; 10: 2197–200. 10.1063/1.1664821 [DOI] [Google Scholar]
- 13. Sutherland B. Quantum many-body problem in one dimension: thermodynamics. J Math Phys 1969; 12: 251–6. 10.1063/1.1665585 [DOI] [Google Scholar]
- 14. Bernard D, Wu YS. A note on statistical interactions and the thermodynamic Bethe ansatz. In: Ge ML and Wu YS (eds.). Proceedings of the 6th Nankai Workshop on “New Developments of Integrable Systems and Long-Range Interaction Models”. Singapore: World Scientific Publishing, 1995.
- 15. Lieb EH, Liniger W. Exact analysis of an interacting Bose gas. I. The general solution and the ground state. Phys Rev 1963; 130: 1605–16. 10.1103/PhysRev.130.1605 [DOI] [Google Scholar]
- 16. Kundu A. Exact solution of double delta function bose gas through an interacting anyon gas. Phys Rev Lett 1999; 83: 1275–8. 10.1103/PhysRevLett.83.1275 [DOI] [Google Scholar]
- 17. Batchelor MT, Guan XW, Oelkers N. One-dimensional interacting anyon gas: low-energy properties and Haldane exclusion statistics. Phys Rev Lett 2006; 96: 210402. 10.1103/PhysRevLett.96.210402 [DOI] [PubMed] [Google Scholar]
- 18. Ha ZNC. Exact dynamical correlation functions of Calogero-Sutherland model and one-dimensional fractional statistics. Phys Rev Lett 1994; 73: 1574–7. 10.1103/PhysRevLett.73.1574 [DOI] [PubMed] [Google Scholar]
- 19. Hashimoto M, He RH, Tanaka Ket al. Particle-hole symmetry breaking in the pseudogap state of Bi2201. Nat Phys 2010; 6: 414–8. 10.1038/nphys1632 [DOI] [Google Scholar]
- 20. Miller TL, Zhang W, Eisaki Het al. Particle-hole asymmetry in the cuprate pseudogap measured with time-resolved spectroscopy. Phys Rev Lett 2017; 118: 097001. 10.1103/PhysRevLett.118.097001 [DOI] [PubMed] [Google Scholar]
- 21. Zhang Y, Wojs A, Jain JK. Landau-level mixing and particle-hole symmetry breaking for spin transitions in the fractional quantum Hall effect. Phys Rev Lett 2016; 117: 116803. 10.1103/PhysRevLett.117.116803 [DOI] [PubMed] [Google Scholar]
- 22. Bhaduri RK, Murthy MVN, Srivastava MK. Fermions at unitarity and Haldane exclusion statistics. J Phys B 2007; 40: 1775–80. 10.1088/0953-4075/40/10/012 [DOI] [Google Scholar]
- 23. Tabert CJ, Carbotte JP. Particle-hole asymmetry in gapped topological insulator surface states. Phys Rev B 2015; 91: 235405. 10.1103/PhysRevB.91.235405 [DOI] [Google Scholar]
- 24. Balakrishnan R, Satija II, Clark CW. Particle-hole asymmetry and brightening of solitons in a strongly repulsive Bose-Einstein condensate. Phys Rev Lett 2009; 103: 230403. 10.1103/PhysRevLett.103.230403 [DOI] [PubMed] [Google Scholar]
- 25. Demchenko DO, Joura AV, Freericks JK. Effect of particle-hole asymmetry on the Mott-Hubbard metal-insulator transition. Phys Rev Lett 2004; 92: 216401. 10.1103/PhysRevLett.92.216401 [DOI] [PubMed] [Google Scholar]
- 26. Kusunose H, Miyake K, Shimizu Yet al. Numerical renormalization-group study of particle-hole symmetry breaking in two-channel Kondo problem: effect of repulsion among conduction electrons and potential scattering. Phys Rev Lett 1996; 76: 271–4. 10.1103/PhysRevLett.76.271 [DOI] [PubMed] [Google Scholar]
- 27. Yang B, Chen YY, Zheng YGet al. Quantum criticality and the Tomonaga-Luttinger liquid in one-dimensional Bose gases. Phys Rev Lett 2017; 119: 165701. 10.1103/PhysRevLett.119.165701 [DOI] [PubMed] [Google Scholar]
- 28. Zhang X, Hung CL, Tung SKet al. Observation of quantum criticality with ultracold atoms in optical lattices. Science 2012; 335: 1070–2. 10.1126/science.1217990 [DOI] [PubMed] [Google Scholar]
- 29. Vogler A, Labouvie R, Stubenrauch Fet al. Thermodynamics of strongly correlated one-dimensional Bose gases. Phys Rev A 2013; 88: 031603(R). 10.1103/PhysRevA.88.031603 [DOI] [Google Scholar]
- 30. Yefsah T, Desbuquois R, Chomaz Let al. Exploring the thermodynamics of a two-dimensional Bose gas. Phys Rev Lett 2011; 107: 130401. 10.1103/PhysRevLett.107.130401 [DOI] [PubMed] [Google Scholar]
- 31. Ha LC, Hung CL, Zhang Xet al. Strongly interacting two-dimensional Bose gases. Phys Rev Lett 2013; 110: 145302. 10.1103/PhysRevLett.110.145302 [DOI] [PubMed] [Google Scholar]
- 32. Hung CL, Zhang X, Gemelke Net al. Observation of scale invariance and universality in two-dimensional Bose gases. Nature 2011; 470: 236–40. 10.1038/nature09722 [DOI] [PubMed] [Google Scholar]
- 33. Batchelor MT, Guan XW. Fermionization and fractional statistics in the strongly interacting one-dimensional Bose gas. Laser Phys Lett 2007; 4: 77–83. 10.1002/lapl.2006100681 [DOI] [Google Scholar]
- 34. Sachdev S, Demler E. Competing orders in thermally fluctuating superconductors in two dimensions. Phys Rev B 2004; 69: 144504. 10.1103/PhysRevB.69.144504 [DOI] [Google Scholar]
- 35. Cazalilla MA, Citro R, Giamarchi Tet al. One dimensional bosons: from condensed matter systems to ultracold gases. Rev Mod Phys 2011; 83: 1405–66. 10.1103/RevModPhys.83.1405 [DOI] [Google Scholar]
- 36. Batchelor MT, Foerster A. Yang-Baxter integrable models in experiments: from condensed matter to ultracold atoms. J Phys A: Math Theor 2016; 49: 173001. 10.1088/1751-8113/49/17/173001 [DOI] [Google Scholar]
- 37. Yang CN, Yang CP. Thermodynamics of a one-dimensional system of bosons with repulsive delta-function interaction. J Math Phys 1969; 10: 1115–22. 10.1063/1.1664947 [DOI] [Google Scholar]
- 38. Takahashi M. Thermodynamics of One-Dimensional Solvable Models. Cambridge: Cambridge University Press, 1999. [Google Scholar]
- 39. Tonks L. The complete equation of state of one, two and three-dimensional gases of hard elastic spheres. Phys Rev 1936; 50: 955–63. 10.1103/PhysRev.50.955 [DOI] [Google Scholar]
- 40. Girardeau M. Relationship between systems of impenetrable bosons and fermions in one dimension. J Math Phys 1960; 1: 516–23. 10.1063/1.1703687 [DOI] [Google Scholar]
- 41. Paredes B, Widera A, Murg Vet al. Tonks-Girardeau gas of ultracold atoms in an optical lattice. Nature 2004; 429: 277–81. 10.1038/nature02530 [DOI] [PubMed] [Google Scholar]
- 42. Kinoshita T, Wenger T, Weiss DS. Observation of a one-dimensional Tonks-Girardeau gas. Science 2004; 305: 1125–8. 10.1126/science.1100700 [DOI] [PubMed] [Google Scholar]
- 43. Bhaduri RK, Murthy MVN, Srivastava MK. Fractional exclusion statistics and two dimensional electron systems. Phys Rev Lett 1996; 76: 165–8. 10.1103/PhysRevLett.76.165 [DOI] [PubMed] [Google Scholar]
- 44. Prokof’ev N, Svistunov B, Tupitsyn I. Worm algorithm in quantum Monte Carlo simulations. Phys Lett A 1998; 238: 253–7. 10.1103/PhysRevE.74.036701 [DOI] [Google Scholar]
- 45. Prokof’ev N, Svistunov B, Tupitsyn I. Exact, complete, and universal continuous-time worldline Monte Carlo approach to the statistics of discrete quantum systems. J Exp Theor Phys 1998; 87: 310–21. 10.1134/1.558661 [DOI] [Google Scholar]
- 46. Rancon A, Dupuis N. Universal thermodynamics of a two-dimensional Bose gas. Phys Rev A 2012; 85: 063607. 10.1103/PhysRevA.85.063607 [DOI] [Google Scholar]
- 47. Nardis JD, Bernard D, Doyon B. Hydrodynamic diffusion in integrable systems. Phys Rev Lett 2018; 121: 160603. 10.1103/PhysRevLett.121.160603 [DOI] [PubMed] [Google Scholar]
- 48. Fletcher RJ, Gaunt AL, Navon Net al. Stability of a unitary Bose gas. Phys Rev Lett 2013; 111: 125303. 10.1103/PhysRevLett.111.125303 [DOI] [PubMed] [Google Scholar]
- 49. Eismann U, Khaykovich L, Laurent Set al. Universal loss dynamics in a unitary Bose gas. Phys Rev X 2016; 6: 021025. 10.1103/PhysRevX.6.021025 [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data supporting the findings of this study are available within the paper and accompanying online supplementary material.












