Anyonic braiding via quench dynamics in fractional quantum Hall liquids
Abstract
In a Laughlin fractional quantum Hall state, one- and two-quasihole states can be obtained by diagonalizing the many-body Hamiltonian with a trapping potential or, for larger systems, from the linear combination of the edge Jack polynomials. The quasihole states live entirely in the subspace of the lowest-energy branch in the energy spectrum with a fixed number of orbits, or a hard-wall confinement. The reduction in the Hilbert space dimension facilitates the study of time evolution of the quasihole states after, say, the removal of the trapping potential. We explore the quench dynamics under a harmonic external potential, which rotates the quasiholes in the droplet, and discuss the effect of long-range interaction and more realistic confinement. Accurate evaluation of the mutual statistics phase of anyons for a wide range of anyon separation can be achieved from the Berry-phase calculation.
pacs:
73.43.Lp, 71.10.PmI Introduction
A remarkable feature of the fractional quantum Hall (FQH) effect Tsui is the emergence of fractionally charged excitations with exotic statistics, dubbed anyons. Laughlin83 ; Wilczek ; Halperin ; Arovas ; WuYS The presence of these collective excitations is a manifestation of the underlying topological phases. Wenbook Abelian anyons are so named because their interchanges lead to an Abelian phase factor , where can be neither an integral multiple of for bosons nor an odd multiple of for fermions. Non-Abelian anyons, on the other hand, live in a degenerate ground-state manifold of topological origin. When they braid around each other, their wave function undergoes a unitary transformation in the degenerate space. MR91 Their potential application as resources in topological quantum computation has aroused great interest in recent years. Kitaev ; NayakRMP
In theory, the property of fractional charge manifests in the Aharonov-Bohm phase acquired by a single anyon along a closed path that enclosed no anyons. The fractional exchange properties, however, arises in the mutual statistical part of the Berry phase, when additional anyons are present in the loop. Arovas So far, experimental evidences mounts for fractional charge of both Abelian (i.e., in a Laughlin-like state) and, presumably, non-Abelian anyons (i.e., in a Moore-Read–like state). Goldman ; Mahalu ; Etienne ; Yacoby ; Mahalu08 ; Radu However, an unambiguous experimental demonstration of fractional statistics has been under intense debate. Inconsistencies between the experimental results and theoretical predictions exist in Fabry-Pérot interferometers Mahalu10 ; West12 ; Umansky ; Willett or Mach-Zehnder interferometers. Ji03 ; Mailly The obstacles in observing the fractional statistics include the Coulomb blockade effect in a small interferometer Halperin11 and the decoherence due to neutral modes in a large interferometer. Umansky19
In novel devices with screening layers conductance oscillations have been observed with abrupt jumps in the interference pattens consistent with the theoretically expected anyonic statistics. Sahasrabudhe ; NakamuraNP19 ; NakamuraNP20 ; WhitNC Moreover, the fractional statistics of FQH quasiholes has also been observed in the current correlation resulting from the collisions among anyons at a beam splitter. BartolomeiScience The achievements heighten the need to study controllable anyon braiding toward the implementation of topological quantum computation. In general, anyons can be created and trapped in the bulk, but dragging them with scanning tunneling microscope (STM) tips is far from reality. On the other hand, in a Fabry-Pérot or Mach-Zehnder interferometer, anyons propagate along sample edges with a finite velocity given by the dispersion of gapless edge modes. Wen1 ; Wen2 ; Wan03 ; Hu09 However, one cannot ignore the details of the edge potential profile in time evolution.
In this work, we construct one- and two-quasihole wave functions in a Laughlin FQH system by introducing an appropriate trapping potential in the interacting Hamiltonian or through linear combinations of Jack polynomials for edge modes. We realize anyonic braiding in the bulk in the quench dynamics after we suddenly remove the tip potential. The projection of the quasihole states to edge wave functions helps us reduce the Hilbert space tremendously in the time evolution. We can introduce an external harmonic potential to confine the anyon excitations and to tune the braiding period, or explore the smear-out of the anyons and their revival in the presence of long-range interaction. We also verify the fractional statistics by calculating the Berry phase in a period, which can be related to the mean angular momentum of the states with various number of quasiholes.
The rest of the paper is organized as follows. In Sec. II we describe how to construct quasihole states with model Hamiltonian or with edge Jack polynomials. We demonstrate in Sec. III the fractional statistics by calculating the mean angular momentum, or equivalently, the Berry phase in the wave-function evolution. In Sec. IV we present a quench protocol to realize the quasihole braiding and discuss the effect of long-range interaction and edge confinement. We conclude in Sec. V and present additional details on constructing quasihole states with Jack polynomials in Appendix A.
II Preparing Quasihole states
Consider a Laughlin FQH droplet with a quasihole state at in a complex plane with model wave function Laughlin83
(1) |
where
(2) |
is the Laughlin wave function with a total angular momentum and is the complex coordinate for the th electron. And is the magnetic length. A two-quasihole wave function can be written as
(3) |
with and being the positions of the quasiholes. In the pseudopotential formalism, HaldanePP is the densest zero energy eigenstate for the -only Hamiltonian. In the interacting Hamiltonian, a trapping potential, which models the effects of an STM tip or a localized impurity, can be introduced to confine a quasihole in the bulk. A previous study Hu08 showed that a repulsive Gaussian potential
(4) |
can be used to trap a quasihole at . Here, and are the strength and width of the trapping potential. In the lowest Landau level with the symmetric gauge, the rotationally invariant basis is
(5) |
and the corresponding matrix elements are . Non-diagonal terms appear when the rotational symmetry is broken by . Without loss of generality, we assume and obtain the integral form of the potential matrix elements as
where and are the confluent hypergeometric function and Bessel function, respectively. In the limit of zero width , the Gaussian can be replaced by a -function , and the matrix elements are simplified to
(6) | |||||
In practice, we use the Gaussian potential with width for the full Coulomb interaction and the -potential for the -only interaction to create a quasihole. Hu08
To simulate the braiding of two quasiholes, we must have a sufficient large system to avoid quasihole overlap. However, large systems are difficult to solve by exact diagonalization due to the absence of rotational symmetry. We, therefore, introduce a truncated subspace for finite systems in disk geometry with the help of Jack polynomials, such that we can calculate up to electrons at filling.

Jack polynomials, or Jacks, are homogeneous symmetric polynomials specified by a rational parameter and a root configuration . Up to the Gaussian factor, FQH model wave functions can often be expressed exactly as Jacks. Bernevig1 ; Bernevig2 We denote the Jack with the root configuration by in the following, while we have a fixed for the Laughlin state and its quasihole and edge excitations. The extra polynomial corresponding to the quasihole excitation can be expanded to be
(7) | |||||
where is the number of electrons. The th term on the right-hand side is a symmetric polynomial of s of order . In general, multiplied by a symmetric polynomial can be interpreted as an edge excitation. LeeJack ; Hu16 For example, and span the edge space with angular momentum . In fact, each term in the above expansion is the unique (zero-energy) ground state of the Hamiltonian in their own momentum subspace for electrons in orbitals, which corresponds to a single Jack. So we can rewrite the one-quasihole state as
(8) | |||||
where the notation represents an unnormalized Jack polynomial and we neglect the Gaussian factor for convenience. In Fig. 1(a), we plot the density profile for a state with one quasihole at on the real axis. The wave functions obtained from diagonalizing the -only Hamiltonian with the -potential at and from the construction by Jacks are identical. To illustrate how this quasihole state can be decomposed into a series of edge states, we plot the low-energy spectrum for the -only Hamiltonian with 10 electrons in 29 orbitals in Fig. 1(b), and use colormap for the zero-energy edge states to indicate the corresponding weight in the one-quasihole state.

To further understand the decomposition of the quasihole state into edge states, we plot the overlap, as defined in Fig. 1(b), as we change the quasihole location in a 3D bar chart in Fig. 2. Here, the momentum of the edge states is defined as . One can see that the edge states with significant weight evolve from having large to zero, when increases from the center to the edge of disk. When is close to half of the radius, the quasihole state incorporates most edge states.

For a two-quasihole state, we consider the following two configurations: (A) with one quasihole at the center and another at , and (B) with two quasiholes symmetrically at . Case A is can be straightforwardly derived from the one-quasihole state with an additional zero inserted to the innermost orbital, i.e.,
(9) | |||||
Therefore, can also be decomposed into a linear combination of the zero-energy eigenstates of the -only Hamiltonian with orbitals and one quasihole fixed at the center of the disk. The density profile of the two-quasihole state with , , and and its overlap with the zero-energy states are shown in Fig. 3. The zero-energy states with non-zero weight in Fig. 3(b) can be regarded as the edge states of the one-quasihole state with the quasihole fixed at the center. The distance between two quasiholes is limited by the radius of the disk. As shown in Fig. 3(a), the second quasihole at can be seen separated from the quasihole at the center and the edge of the droplet.
For the same system size, the distance of the quasiholes can be further separated in case B as shown in Fig. 4(a). However, the construction of the two-quasihole state via Jacks is more complicated. The model wave function with the two quasiholes at is
(10) | |||||
the polynomial factor of which can be expanded to be
(11) | |||||
Following the same procedure in the one-quasihole case, we find that the involved edge states are all zero-energy states for the system with orbitals and a total angular momentum . The symmetrical arrangement of the quasiholes further rules out the contribution from the states with odd . As in the one-quasihole case, the increase of leads to increasing weight in the small-momentum subspace. However, each zero-energy momentum subspace can now be degenerate. The degeneracy prevents us to construct the wave function by a single Jack in each subspace. For electrons, e.g., the degeneracies are for , as shown in Fig. 4. Those degeneracies could be obtained by counting the number of root configurations of the Jacks for electrons in orbitals. In particular, the degeneracy is the largest at and is the same for and . In order to expand by edge states in a large system, we need to know how to map the polynomials in Eq. (11) to linear combinations of edge Jacks, which we elaborate in Appendix A.

III fractional statistics


With the exact one-quasihole and two-quasihole wave functions, we can verify the fractional statistics via the Berry phase calculation when the system is adiabatically rotated by a full circle. In Eqs. (8)-(10), the rotation is achieved by letting and varying in the wave functions , such that the Berry phase is defined as
(12) |
The rotation could alternatively be generated by the angular momentum component as
(13) | |||||
Therefore, we have
(14) |
In our setup, therefore, the mean angular momentum can be related to the fractional statistics of the quasihole. Equivalently, one can calculate the mean square radius for the density distribution, because the radius of the th Landau orbital (with angular momentum ) is simply in the disk geometry. Carusotto18 ; Carusotto19 ; Lewenstein20
As a result, the mutual braiding statistical phase of two quasiholes can be expressed by the Berry phase of the two-quasihole state subtracting the Berry phases for the two single-quasihole states,
(15) |
where and are the mean angular momentum for the one-quasihole states with a quasihole at and 0 for case A, or for case B. The total angular momentum of the Laughlin state is included in Eq. (15) to compensate for the contributions from the constituent electrons. In Fig. 5, we plot the statistical phase as a function of the distance of quasiholes for the two cases. The results are obtained from numerical diagonalizing the model Hamiltonian with -function impurity potential for electrons and from the Jack polynomial contruction for electrons. The expected value of for a Laughlin quasihole can be anticipated when (1) the two quasiholes are not too close to each other, and (2) neither quasihole is close to the edge of the droplet. The plateau of is well developed in case B, in which the distance between quasiholes doubles that in case A, while the influence of the edge is the same. The plateau starts from a quasihole distance of about , which is approximately the diameter of a Laughlin quasihole. WuYL This implies that quasiholes can be thought of as independent of each other, as long as they are not overlapping with each other. Fig. 5 further suggests that when the two quasiholes overlap with each other, their interaction, as reflected in the correction to , is independent of the system size, hence local.
IV The quench evolution of the two-quasihole state
In previous sections we have demonstrated that one- and two-quasihole states of a Laughlin droplet can be constructed by Jack polynomials, which allows us to set up larger systems with two anyons that are sufficiently far from each other and from the edge, such that their mutual fractional statistics can be computed. In this section we explore the dynamical evolution of the quasihole states after a quantum quench. The exploration of such a quench protocol is motivated by the implementation of topological quantum computing, which needs the braiding of anyons. In our setup the system starts with quasiholes localized in a properly designed trapping potential, which is removed or altered at a certain time, say, . In the ideal case one quasihole can move around another while still maintaining their local density profile. This scheme may be advantageous in realizing braiding, e.g., in GaAs-based systems, in which the two-dimensional electron gas is far from the surface gates. It might also be the preferred protocol once fractional quantum Hall systems are realized in cold atomic Cooper ; Gemelke ; Miyake ; Aide ; He ; Tobias or photonic LLphoton ; photonFQH ; Zhou systems in synthetic gauge field.
In the first realization we show that the rotation of a quasihole around the other can be realized by a quench protocol, in which a harmonic potential can introduce an angular velocity. The harmonic potential, in the form of in real space, realizes a linear potential in the momentum space, which introduces a linear dispersion for the edge states. We start, in the protocol, with a Hamiltonian for the Laughlin liquid and two tip potentials that trap two quasiholes at . In the edge Jack polynomial expansion of the quasihole state as in Eq.(11), each term has a fixed angular momentum, which is inversely linear in the power of . At , we turn off the tip potentials and, instead, turn on the harmonic potential with . After the quench, the system evolves as , in which the reduced Hamiltonian contains the potential only. Therefore, the dynamical phase difference of these terms can be absorbed by the corresponding polynomials of , leading to a pure rotation. Fig. 6 shows the evolution of the density profile of and the overlap of the wave function with for electrons in orbitals. We find that the overlap shows regular oscillations, and the density profile is stable in time. The period of the oscillations is , which corresponds to the exchange of two quasiholes, for which the wave function picks up a negative sign only.

The minima in the time-dependent overlap is wider than the maxima in Fig. 6. This has an interesting origin and can be used to estimate the size of quasiholes. As we have discussed, the time evolution of the wave function is known analytically:
(16) |
where . In Fig. 7(a) we plot the overlap in a period with different for a system with electrons in orbitals. For a small , the oscillation is sinusoidal, and a nonzero minimum appears at (indicated by the vertical arrow), when the quasiholes are rotated to . As increases, the overlap at decreases and vanishes at , whose value is close to the radius of a quasihole , which indicates that the nonzero value of the minimum may come from the spatial overlap of the quasiholes. Further, the overlap deviates from zero, again, at , presumably when the quasiholes touch the rim of the droplet. The continuous dependence of the overlap at on is shown in Fig. 7(b). To understand the curves in Fig. 7 and their connection to the quasihole size, we model quasiholes by rigid disk of radius . As illustrated in the inset of Fig. 7(b), if the wave function overlap vanishes when quasiholes no longer overlap with each other (for sufficiently large ), there exists a characteristic angle of rotation
(17) |
In Fig. 7(a) we mark the angles for by red dots on the corresponding curves and find that they explain well the flat bottom of the overlap curves.

A more realistic situation can involve some long-range interaction and a neutralizing confining potential. We consider the following mixed Hamiltonian:
(18) |
where is the model Hamiltonian and is the Coulomb Hamiltonian
(19) |
which contains the Coulomb interaction among electrons and between electrons and background charges Wan03 . Here are parameters that we tune to adjust the bandwidth of the low-energy edge-state manifold and its gap to the higher-energy states. As in Fig. 4, we choose and . In general, the direct calculation of the matrix exponential in is time and memory consuming for large systems. For example, the two-quasihole state for -electron has a full dimension of 1,251,677,700 in orbitals. The situation is significantly simpler when we assume that the quasiholes are prepared in the model wave function in Eq. (10), which can be written as the linear combination of a few edge states, as demonstrated in Fig. 4. These edge states are protected by an energy gap, which origins from a hard wall edge potential due to the choice of the fixed number of orbitals. If we assume that the quasihole state do not leak outside the low-energy manifold, the time evolution can be calculated only in the truncated edge space. For the -electron two-quasihole state, the dimension of the reduced Hamiltonian is merely and we have
(20) |
where and are the energy and wave function of the relevant edge states, and we have . However, due to the presence of Coulomb interaction, the dispersion of the edge modes is no longer linear, as is evident in Fig. 4. In Fig. 8 we plot the wave function overlap as a function of time. The curve oscillates with multiple periods. In particular, we can identify a period of , at the integral multiple of which the overlap is about 0.75 or more. A smaller period, which is about is also visible in the overlap. Meanwhile, the plots of density profiles indicate that the nonlinear dispersion of the edge mode deforms the shape of the quasihole in a long time evolution. The smear-out of the quasiholes is significant when the overlap is small.


The oscillations in the overlap can be understood as follows. According to Fig. 4, the expansion of the time-dependent wave function in Eq. (20) is dominated by a few edge states (only six states have a weight larger than 0.01). We can calculate the overlap to be
(21) |
Therefore, the time-dependent overlap is dominated by the energy differences of two edge states, both with significant probability amplitudes. To show this, we calculate the Fourier transformation of the data in Fig. 8 and plot the spectral weight in Fig. 9. In the inset we copy the part of the low-energy spectrum in Fig. 4(b) that includes the edge states with appreciable weight. In particular, we label the six most important components by , according to the order of their weights. The peaks in the Fourier spectrum are in good agreement with the energy differences among them, as listed in Table 1. Specifically, we have , which corresponds to a period ; this explains the period of the major revivals of the overlap. The subleading peak corresponds to , which is 7.2 times larger than ; this explains the shorter period that appears in Fig. 8. The third leading peak is located at ; the corresponding period is not easy to detect by visual examination of Fig. 8 as it is close to and roughly . The leading seven contributions, resulting in the highest peaks in Fig. 9 (notice ), are underlined in Table 1.
V Discussions and Conclusions
In summary, we construct one-quasihole and two-quasihole wave functions for a Laughlin quantum Hall state via numerically diagonalizing the model Hamiltonian with an impurity potential or the combination of the edge Jacks. In the Jack polynomial construction, we identify a universal mapping between the edge Jacks and the Jastrow term of the Laughlin wave function multiplied by symmetric polynomials, which allows us to approach systems with up to electrons. These quasihole states lives entirely in the degenerate zero-energy Hilbert space with a hard-wall confinement; this fact tremendously reduces the dimension of the Hilbert space in the time evolution.
The fractional statistics of the quasiholes can be computed through the Berry phase or, effectively, the mean angular momentum of those states, as long as the quasiholes are not overlapping with each other and with the edge of the FQH droplet. The novel technique with Jack polynomials provides us a reasonably wide range of quasihole location, in which we can observe a plateau of the expected anyon statistics.
Based on a quench protocol, we demonstrate that quasiholes in the bulk can be driven by a harmonic potential or, in the more realistic case, by Coulomb interaction and background charge confinement to braid around each other. This reveals interesting possibilities of manipulating anyons by dynamical quench in numerical simulations and, possibly, in future experiments toward the realization of topological quantum computation with anyons in FQH liquids.
0 | ||||||
0.00655 | 0 | |||||
0.04717 | 0.05373 | 0 | ||||
0.02016 | 0.02672 | 0.02701 | 0 | |||
0.00387 | 0.01042 | 0.04330 | 0.01629 | 0 | ||
0.01860 | 0.02516 | 0.02857 | 0.00156 | 0.01473 | 0 |
Acknowledgements.
This work at Chongqing University was supported by National Natural Science Foundation of China No.11974064, the Chongqing Research Program of Basic Research and Frontier Technology No.cstc2021jcyj-msxmX0081, Chongqing Talents: Exceptional Young Talents Project No.cstc2021ycjh-bgzxm0147, and the Fundamental Research Funds for the Central Universities No. 2020CDJQY-Z003. N.J. was supported by National Natural Science Foundation of China No.12104075. X.W. was supported by the Strategic Priority Research Program of Chinese Academy of Sciences through Grant No. XDB28000000.Appendix A Mapping Two-Quasihole States to Jack Polynomials
In Sec. II we discussed the principles and difficulties of constructing a two-quasihole state with a symmetric configuration as in case B. The apparent difficulty is that each term in Eq. (11), the polynomial expansion of the model wave function, involves multiple zero-energy Jacks in the same momentum subspace. We need to find a series of matrices in these subspaces that can map the polynomials to the corresponding Jacks.
For simplicity, let us take a -electron Laughlin state as a concrete example. Multiplied by the Jastrow term , the first term on the right-hand side of Eq. (11) corresponds uniquely to the Jack , which lies in the momentum subspace with . The second term , however, corresponds to the Jack and its descendant from the squeeze rule . They both have and span a two-dimensional degenerate subspace. In the same manner, the third term in Eq. (11) can be expanded by and its two descendants and , which result in a three-fold degeneracy in the subspace. Now the question is how to combine degenerate Jacks to construct all the terms in Eq. (11) multiplied by the Jastrow factor. In other words, we need to calculate the coefficient in the expression
(22) |
where is the polynomial in Eq. (11) with degree . The summation of contains all the Jacks whose root configurations descend from , which we denote for the Jack with electrons (or 1s) to the right of the two inserted quasiholes (or 0s, labelled in red). The coefficients need to be determined with the nonorthogonality among and its descendants in mind. The details resemble the monomial expansion of an arbitrary monomial multiplied by the Laughlin state, which we need in the construction of edge excitations; in the latter case, however, the summation is taken without restricting the orbital number. LeeJack Therefore, we have here a simpler situation in which the two-quasihole state lives in fixed orbitals, so the degeneracy is smaller. Taking the 4-electron system as an example, we determine for by solving the following identity
where is the Laughlin state and is a lower triangular matrix in the following form
(23) |
For , we need to solve
which yields a lower triangle matrix
(24) |
For , we need, again, a matrix, which turns out to be identical to and satisfies
These are the matrices needed for the 4-electron system with 12 orbitals, whose degeneracies in the zero-energy subspace for are , respectively. But in a system with more than 5 electrons, the degeneracy for becomes 4, and the corresponding matrix is
(25) |
Interestingly, the -dimensional matrix is embedded as the lower right submatrix of the -dimensional matrix . We find these matrices are independent of the system size and, as shown in the 4-electron case, the same for and . Since the largest system we calculate in this work contains 12 electrons, all we need to present is the following matrix for subspace:
(26) |
When two quasiholes are arranged in an asymmetric configuration, i.e., , the odd subspaces are also involved. The matrices for these momentum space are, again, symmetric around . It still holds that is embedded as the lower right submatrix in . So, we only need to present the largest matrix for and :
(27) |
Finally, we emphasize that the lowest row in each matrix corresponds to the coefficients in Eq. (22). Therefore, we can construct the product of each term in Eq. (11) and the Jastrow term by a linear combination of Jack polynomials. This concludes our construction of the two-quasihole wave function with only matrix, or in the more general asymmetric case with both and matrix.
References
- (1) D. C. Tsui, H. L. Störmer, and A. C. Gossard, Phys. Rev. Lett. 48, 1559 (1982).
- (2) R. B. Laughlin, Phys. Rev. Lett. 50, 1395 (1983).
- (3) F. Wilczek, Phys. Rev. Lett. 49, 957 (1982).
- (4) B. I. Halperin, Phys. Rev. Lett. 52, 1583 (1984).
- (5) D. Arovas, J. R. Schrieffer, and F. Wilczek, Phys. Rev. Lett. 53, 722 (1984).
- (6) Y. S. Wu, Phys. Rev. Lett. 52, 2103 (1984).
- (7) X.-G. Wen, Quantum Field Theory of Many-Body Systems (Oxford University Press, Oxford, 2004).
- (8) G. Moore and N. Read, Nucl. Phys. B 360, 362 (1991).
- (9) A. Y. Kitaev, Ann. Phys. 303, 2 (2003).
- (10) C. Nayak, S. H. Simon, A. Stern, M. Freedom, and S. Das Sarma, Rev. Mod. Phys. 80, 1083 (2008).
- (11) V. J. Goldman and B. Su, Science 267, 1010 (1995).
- (12) R. De-Picciotto, M. Reznikov, M. Heiblum, V. Umansky, G. Bunin, and D. Mahalu, Nature 389, 162 (1997).
- (13) L. Saminadayar, D. C. Glattli, Y. Jin, and B. Etienne, Phys. Rev. Lett. 79, 2526 (1997).
- (14) J. Martin, S. Ilani, B. Verdene, J. Smet, V. Umansky, D. Mahalu, D. Schuh, G. Abstreiter, and A. Yacoby, Science 305, 980 (2004).
- (15) M. Dolev, M. Heiblum, V. Umansky, A. Stern, and D. Mahalu, Nature 452, 829 (2008).
- (16) I. Radu, J. B. Miller, C. M. Marcus, M. A. Kastner, L. N. Pfeiffer, and K. W. West, Science 320, 899 (2008).
- (17) N. Ofek, A. Bid, M, Heiblum, A. Stern, V. Umansky, and D. Mahalu, Proc. Natl. Acad. Sci. 107, 5276 (2010).
- (18) D. T. McClure, W. Chang, C. M. Marcus, L. N. Pfeiffer, and K. W. West, Phys. Rev. Lett. 108, 256804 (2012).
- (19) I. Sivan, R. Bhattacharyya, H. K. Choi, M. Heiblum, D. E. Feldman, D. Mahalu, and V. Umansky, Phys. Rev. B 97, 125405 (2018).
- (20) R. L. Willett, L. N. Pfeiffer, and K. W. West, Proc. Natl. Acad. Sci. 106, 8853 (2009).
- (21) Y. Ji, Y. Chung, D. Sprinzak, M. Heiblum, D. Mahalu, and H. Shtrikman, Nature 422, 415 (2003).
- (22) P. Roulleau, F. Portier, D. C. Glattli, P. Roche, A. Cavanna, G. Faini, U.Gennser, and D. Mailly, Phys. Rev. B 76, 161309 (2007).
- (23) B. I. Halperin, A. Stern, I. Neder, and B. Rosenow, Phys. Rev. B 83, 155440 (2011).
- (24) R. Bhattacharya, M. Banerjee, M. Heiblum, D. Mahalu, and V. Umansky, Phys. Rev. Lett. 122, 246801 (2019).
- (25) H. Sahasrabudhe et al., Phys. Rev. B 97, 085302 (2018).
- (26) J. Nakamura, S. Fallahi, H. Sahasrabudhe, R. Rahman, S. Liang, G. C. Gardner, and M. J. Manfra, Nat. Phys. 15, 563 (2019).
- (27) J. Nakamura, S. Liang, G. C. Gardner, and M. J. Manfra, Nat. Phys. 16, 931 (2020).
- (28) A. M. Whiticar, A. Fornieri, E. C. T. ÓFarrell, A. C. C. Drachmann, T. Wang, C. Thomas, S. Gronin, R. Kallaher, G. C. Gardner, M. J. Manfra, C. M. Marcus, and F. Nichele, Nat. Comm. 11, 3212 (2020).
- (29) H. Bartolomei, M. Kumar, R. Bisognin, A. Marguerite, J. M. Berrior, E. Bocquillon, B. Placais, A. Cavanna, Q. Dong, U. Gennser, Y. Jin, and G. Feve, Science 368, 173 (2020).
- (30) X. G. Wen, Int. J. Mod. Phys. B 6, 1711 (1992).
- (31) X. G. Wen, Phys. Rev. B 41, 12838 (1990).
- (32) X. Wan, E. H. Rezayi, and K. Yang, Phys. Rev. B 68, 125307 (2003).
- (33) Z. X. Hu, E. H. Rezayi, X. Wan, and K. Yang, Phys. Rev. B 80, 235330 (2009).
- (34) F. D. M. Haldane, Phys. Rev. Lett. 51, 605 (1983).
- (35) Z. X. Hu, X. Wan, and P. Schmitteckert, Phys. Rev. B 77, 075331 (2008).
- (36) B. A. Bernevig and F. D. M. Haldane, Phys. Rev. Lett. 100, 246802 (2008).
- (37) B. A. Bernevig and F. D. M. Haldane, Phys. Rev. Lett. 101, 246806 (2008).
- (38) K. H. Lee, Z. X. Hu, and X. Wan, Phys. Rev. B 89, 165124 (2014).
- (39) N. Jiang and Z. X. Hu, Phys. Rev. B 94, 125116(2016).
- (40) R. O. Umucalılar, E. Macaluso, T. Comparin, and I. Carusotto, Phys. Rev. Lett. 120, 230403 (2018).
- (41) E. Macaluso, T. Comparin, L. Mazza, and I. Carusotto, Phys. Rev. Lett. 123, 266801 (2019).
- (42) T. Gra, B. Juliá-Díaz, N. Baldelli, U. Bhattacharya, and M. Lewenstein, Phys. Rev. Lett. 125, 136801 (2020).
- (43) Y.-L. Wu, B. Estienne, N. Regnault, and B. A. Bernevig, Phys. Rev. Lett. 113, 116801 (2014).
- (44) N. R. Cooper, Adv. Phys. 57, 539 (2008).
- (45) N. Gemelke, E. Sarajlic, and S. Chu, arXiv:1007.2677.
- (46) H. Miyake, G. A. Siviloglou, C. J. Kennedy, W. C. Burton, and W. Ketterle, Phys. Rev. Lett. 111, 185302 (2013).
- (47) M. Aidelsburger, M. Atala, M. Lohse, J. T. Barreiro, B. Paredes, and I. Bloch, Phys. Rev. Lett. 111, 185301 (2013).
- (48) Y.-C. He, F. Grusdt, A. Kaufman, M. Greiner, and A. Vishwanath, Phys. Rev. B. 96, 201103 (2017).
- (49) B. Andrade, V. Kasper, M. Lewenstein, C. Weitenberg, and T. Graß Phys. Rev. A 103, 063325 (2021).
- (50) N. Schine, A. Ryou, A. Gromov, A. Sommer, and J. Simon, Nature 534, 671 (2016).
- (51) L. W. Clark, N. Schine, C. Baum, N. Jia, and J. Simon, Nature 582, 41 (2020).
- (52) Y. Yan, Q. Zhou, arXiv:2106.15726.