The theory of topological quantum computation is underpinned by two important classes of models. One is based on non-abelian Chern–Simons theory, which yields the so-called anyon models that often appear in the context of electrically charged quantum fluids. The physics of the other is captured by symmetry broken Yang–Mills theory in the absence of a Chern–Simons term and results in the so-called quantum double models. Extensive resources have been invested into the search for anyon quasi-particles, in particular, the so-called Ising anyons (k = 2) of which Majorana zero modes are believed to be an incarnation. In contrast to the models, quantum doubles have attracted little attention in experiments despite their pivotal role in the theory of error correction. Beyond topological error correcting codes, the appearance of quantum doubles has been limited to contexts primarily within mathematical physics, and as such, they are of seemingly little relevance for the study of experimentally tangible systems. However, recent works suggest that quantum double anyons may be found in spinor Bose–Einstein condensates. In light of this, the core purpose of this article is to provide a self-contained exposition of the quantum double structure, framed in the context of spinor condensates, by constructing explicitly the quantum doubles for various ground state symmetry groups and discuss their experimental realisability. We also derive analytically an equation for the quantum double Clebsch–Gordan coefficients from which the relevant braid matrices can be worked out. Finally, the existence of a particle-vortex duality is exposed and illuminated upon in this context.
I. INTRODUCTION
As one ventures beyond the realm of classical phases of matter, the classification paradigm due to Landau1 ceases to apply. Such quantum systems often exhibit phenomena of a long-range collective nature connected to a non-trivial underlying topological structure. Consequently, topology may serve as a better fingerprint to categorize such phases. Owing to its inherent relationship to the connectivity of space, the spin-statistics theorem generally breaks down on non-trivial topologies, thus clearing the way for more exotic particle species to emerge that may not be classified as bosons or fermions. Such peculiar particles exhibit fractional statistics and are, therefore, referred to as anyons2–5 (any as in any statistics). Permutation of anyons will, in contrast to fermions and bosons, generally implement a more complex unitary transformation than a simple change of sign of the wavefunction.
While anyons are very interesting from a fundamental viewpoint, most research on the topic concerns quantum information processing.6,7 A strong interest in anyons was sparked when Kitaev suggested that they may hold the key to the realization of fault tolerant quantum computation.8 The prospects for quantum computation are tantalizing, but in order to not succumb to decoherence, the effects of environmental noise must be considered. This is naturally addressed in a quantum computer based on anyons. Owing to the intrinsic properties of topology, qubit states based on anyons possess a natural shield against various unwanted interactions with the environment. Error-immune computers of this kind are known as topological quantum computers (TQCs).8–12
Anyons may be realized in planar fermionic fluids, such as an electron gas in the fractional quantum Hall effect,13 which effectively are governed by the Chern–Simons theory.14–16 The Chern–Simons term is responsible for a type of charge–flux attachment, akin to that in the Aharanov–Bohm experiment,17,18 which further results in quasi-particles with fractional statistics. These models are known as models,19,20 where the parameter is an integer referring to a particular deformed representation of . A strong interest for Majorana quasi-particle zero modes has been developed in recent years since they are believed to realize the Ising anyons.20–22 Majorana zero modes have been predicted to emerge in the vortex quasi-particle spectrum of chiral p-wave superfluids such as in 3He superfluids or cold atom Fermi gases.23–27 They are also expected to emerge in solid state systems such as in fractional quantum Hall fluids with filling fraction ,28–30 in superconducting–semiconducting nanowires,31–38 and in the vortex cores of certain topological superconductors.39
Moreover, anyons are also believed to emerge in certain bosonic systems, such as Bose–Einstein condensates (BECs),40 as a result of spontaneous breaking of the initial continuous gauge symmetry to a discrete residual subgroup. Models based on spontaneously broken gauge symmetries are known as quantum doubles.8,41–44 A quantum double can be viewed as an emergent low-temperature symmetry algebra, where the group structure is “doubled” by combining it with its Fourier dual. Mathematically they constitute, just like the models, examples of so-called quantum groups.43,45 In particular, quantum group symmetry emerges when the degrees of freedom pertaining to (generalized) electric charges interact with those of (generalized) magnetic fluxes. Consequently, the particle content of an anyon model is labeled by the irreducible representations of the pertinent quantum group. This is in contrast to conventional quantum field theory, where the particles are labeled by those of an undeformed group. For the sake of completeness, a detailed exposition of the particular quantum group relevant to this work is provided in Appendix C and is also thoroughly discussed in Ref. 41.
The quantum double structure has been studied extensively from a mathematical perspective, yet little efforts have been made to reconcile it with real physical systems. It is not well known that low-temperature phases of spinor BECs46–52 with discrete residual symmetry are underpinned by a quantum double structure. Hence, this work primarily seeks to introduce the concept of quantum doubles to the cold atoms community and to provide concrete examples demonstrating how the various components of the quantum double structure might materialize in spinor BECs. There are mainly two types of quantum double excitations referred to as fluxons and chargeons. While the fluxons, as per homotopy theory, correspond to quantized vortices, the physical interpretation of chargeons is not that clear. By illustrating that fluxons and chargeons are in fact Fourier duals of one another, we find that the chargeons may be associated with delocalized waves. In particular, by conducting an analysis of the normal modes, the chargeons appear to be represented as spin rotations and spin waves, also known as magnons.53–56 This agrees with results obtained from numerical simulations carried out in Ref. 57, where spin waves were observed as remnants after fusion events involving non-abelian vortices. The fact that spinor BECs can be routinely produced in the laboratories58–62 brings further justification supporting the study of non-abelian anyons in such systems. Before moving on, we wish to highlight that the field theory descriptions of anyons in fermionic and bosonic systems are not always of a Chern–Simons and Yang–Mills type, respectively. Indeed, there exist systems that are based on spin-1/2 particles, which are underpinned by quantum doubles, for example, Kitaev's toric code8 and honeycomb model.63 Similarly, there are systems whose constituent particles are bosons and are described by Chern–Simons theory. A Chern–Simons gauge gravity is an example of the latter.64,65
II. CHERN–SIMONS ANYONS
Before discussing the quantum double structure in spinor BECs, we describe anyons in fermionic systems of a Chern–Simons type for the purpose of providing a more cohesive perspective on anyons. In addition, the classification scheme for topological defects is described, which is later applied to spinor BECs in Secs. III E and III F to derive their topological charge.
A. From Aharonov–Bohm effect to topological quantum field theory
Let us consider the Aharonov–Bohm setup.17 In this experiment, an electron encircles an (effectively) infinitely long tube of magnetic flux, see Fig. 1, which is piercing through three-dimensional (3D) space. As the electron traverses around the flux-tube, the wave-function accumulates a complex phase due to the non-zero vector (gauge) potential originating from the flux.
1. Charge–flux attachment
B. Symmetry classification of topological particles
The emergent charge–flux attachment concept discussed in Sec. II A is unique to (2 + 1)-dimensions. Imagine then that there are n flux tubes enclosed by a loop. The loop may be continuously deformed such that it can be decomposed into n distinct loops, each of which encircles one tube of magnetic flux [see Figs. 3(d)–3(f)]. This means that the total flux αB calculated in Eq. (4) now is . The integer n counts the phase winding, which is why it is known as the winding number, and as illustrated in Fig. 3, it must be a topological invariant. Similarly, we illustrate in Figs. 3(a)–3(c) how loops can be attached, which allows for the combined flux to be deduced.
The group of integers possesses abelian structure, which is why the flux–charge composites considered here are abelian anyons. However, we may also consider more general gauge theories in which the topology of the gauge group is more complicated and for a gauge group G, the fundamental group need not be isomorphic to the group of integers. Contrary to bosons and fermions, if two flux-charge composites in such a theory were to be exchanged twice, the wave function may not return to its original state as an arbitrary phase factor may be acquired.
C. Ising anyon topological quantum computer
However, this model is computationally non-universal, meaning that the braid group generated from σ1 and σ2 is finite and is, thus, only capable of implementing a finite set of unitary rotations of the Bloch sphere. In order to be able to implement every logic gate, we need a braid group of infinite order that is generating a topologically dense cover in , so that any rotation of the Bloch sphere can be realized. Nevertheless, universality can be achieved by supplementing the set of braids with an additional conventional operation at the cost of sacrificing the complete topological protection.81,82 Such a quantum computer is, therefore, to be regarded as a hybrid as arbitrary computational processes will still rely on error correcting protocols to some extent. Next, we shall redirect our attention toward bosonic systems and their quantum double structure in phases with spontaneously broken symmetry.
III. QUANTUM DOUBLE ANYONS
In Sec. II A, we described how two-dimensional electrodynamics with interactions governed by continuous gauge symmetry give rise to an Aharonov–Bohm like phase. The emergence of this phase is due to the punctures caused by the vortex-like excitations (fluxons), which, thus, play the role of flux tubes. We also saw that a modified Gauss's law emerged as a consequence of the additional Chern–Simons term, which lead to a flux-charge attachment, and thus, to fractional statistics. Topological physics can also be realized in the absence of such a term via the Higg's mechanism, since the gauge bosons of the theory may acquire mass, thus resulting in a topological field theory, as the non-topological interactions are rendered short range.
A. The general picture
1. Discrete gauge theory in (2 + 1)-D
Here, we shall consider bosonic condensates which have undergone spontaneous symmetry breaking from a continuous symmetry group G to a finite discrete subgroup H, where each subgroup H serves as a signature of a low-temperature phase. We refer to this group as the isotropy group of the condensate as it does not alter the physical properties of the system. However, since all cosets of the isotropy group possess the same structure, the full order parameter manifold is defined by the quotient space . Just like in Sec. II B, where it was described how the fluxons can be classified by means of the first homotopy group over the order parameter manifold , here, the excitations are classified according to . However, if H is discrete and G/H is simply connected, they are connected by an isomorphism so , which entails that all information about the excitations is encoded in H, so that the elements of H may be used for labeling the excitations. However, the exact classification is slightly more subtle since charge conservation must be respected. That is, if we regard two fluxons and bring the hi fluxon around the hj one, it returns in a transformed state . Flux conservation, thus, enforces the condition , which further entails that the state must transform under conjugation . Conjugation defines an equivalence relation on the group so the fluxons should, therefore, be classified according to the conjugacy class (CC) partitioning, which consequently implies that all particles labeled by an element within the same conjugacy class are indistinguishable. This phenomenon is known as flux metamorphosis,83,84 and if the residual group H is non-commutative, there are generally more than one element in each conjugacy class which together span a multi-dimensional Hilbert space inhabited by a non-abelian fluxon.
2. Higg's mechanism and symmetry breaking
Inheriting the nomenclature adopted in the Aharonov–Bohm experiment, the fluxons play the role of flux tubes while the matter fields correspond to electric charges, or chargeons, which are arranged according . If a chargeon labeled by some UIR encircles a fluxon labeled by some element hj the Aharonov–Bohm phase is given by . Thus, due to the minimal coupling introduced in the covariant derivative in Eq. (21), we have a notion of charge–flux attachment akin to that in Eq. (10). Such composite objects are called dyons. However, the labeling of the dyons is subtle and requires a more careful analysis. Following Ref. 85, we shall consider the Aharonov–Bohm experiment with two dyons where one of them is hidden between two slits in a plate placed in front of a screen. The flux part of the two dyons is labeled by hi and hj, respectively, so if the first dyon goes through the left slit, the flux part of the dyon between the slits transforms according to . However, if the first dyon goes through the right slit, the second one sitting between the slits is left invariant, i.e., , and consequently, we have asymmetry between the two beams. This is due to the fact that there is a Dirac string connecting the fluxon–anti-fluxon pairs, which is crossed only when passing the fluxon on one of its sides (see Fig. 6). Owing to this correlation, we only have constructive interference if hi and hj commute since the hi will slip through so that . This implies that the charge attached to hj must transform under UIRs of elements that commute with hj. Such a set of elements always possesses group structure and is known as the centralizer group of hj. In conclusion, the adequate labeling of the dyons is given by the conjugacy classes Ci partitioning H (the fluxon part) and the UIRs of the centralizers of the conjugacy classes (the chargeon part). We have now dissected the structure of the quantum double construction, which revealed that fluxon Hilbert spaces are spanned by the elements of the conjugacy class considered. That is, a generic fluxon state is a coherent superposition , where hi are elements within the same conjugacy class and ci are coefficients belonging to the field . Moreover, the chargeon Hilbert space is a space of functions on H (the centralizer UIRs), which means that the full quantum double Hilbert space is given by their tensor product . Before we move on and discuss the quantum double algebra in more detail, we shall provide an example of a simple quantum double model based on the group , which will help us to motivate the various components of the structure.
B. lattice electrodynamics—A simple example
We can also derive the monodromy by creating string operators that, for instance, are bringing an e around an m, as in Fig. 8(b). Note that the loop corresponds to the same operation as the product of the plaquettes it encircles. All of the plaquettes return eigenvalue 1 except the one on which the fluxon is residing. That is, . The eigenvalue of a braiding process where an e is braided twice with an m is, thus, given by ω. This is, in fact, a discrete version of Stokes' theorem where the plaquette inhabited by the fluxon carries circulation so that the loop can be smoothly deformed around this plaquette, without affecting the outcome. As shown in Fig. 8(b), the Dirac string connecting m and is crossed once independent of the deformation of the loop, which results in a phase ω due to the relation between X and Y in Eq. (24). We may, thus, conclude that topological equivalence is naturally encoded in this model. The plaquette operator acting on the fluxon can be regarded as a measurement operator that is projecting out the flux. In electromagnetism, this is equivalent to and the action of the vertex operator, thus, represents something akin to Gauss's law . Generally, the flux is measured by Z operators while the X operators correspond to gauge transformations. Hence, the dyons, which can be regarded as the elementary objects of the quantum double structure, must be arranged according to the UIRs implementing the action of a flux measurement followed by a gauge transformation.
1. Chargeon–fluxon duality
C. The quantum double construction
D. Vortex chromodynamics picture
E. Anyons in spin-0 and spin-1 BECs
In a spinor BEC, the order parameter representing the phase under consideration is not a scalar but a spinor. The order parameter wave function of a spin-F condensate belongs to the complex vector space , which is isomorphic to the real space , i.e., , where S is the spin. This entails that the order parameter is a map onto a surface isomorphic to a sphere, that is, it maps onto the manifold . Consequently, we can conclude that the full symmetry group of the system is accompanied by phase invariance, i.e., . Here, we consider two types of condensates with spin-0 and spin-1 degrees of freedom, respectively, which are both governed by abelian theories. In the S = 0 case, we know that the corresponding wave function is a scalar . Moreover, if we break the symmetry, we know that each subgroup is inheriting the commutative property, more precisely each subgroup must be isomorphic to a group, thus giving rise to a quantum double akin to the lattice model discussed in Sec. III B. In the spin-1 case, we have three low-temperature inert states , and , which correspond to the two ferromagnetic states and the polar state, respectively. The symmetry groups of these states are given by rotations about the axes parallel to the states so that reduces to for the ferromagnetic states and for the polar state. Here, denotes the infinite dihedral group which also includes reflection symmetry, as opposed to . As for discrete symmetries in these phases, we can conclude that we only have one corresponding to the reflection part of the symmetry, which can be represented by the cyclic group C2.
A convenient way to deduce the symmetries is to consider a graphical representations of the spinorial states, such as the spherical harmonics or the Majorana star representation.47 Note that since these groups are abelian, spin-0 or a spin-1 condensates are not able to support non-abelian anyons as their natural excitations. This fact can be established from the analysis outlined in Sec. II B, where we discussed how the topological excitations in planar systems can be classified according to the first homotopy group of the coset space G/H generated by the subgroup H.
1. The quantum double of C2
2. Particle-vortex duality in
Interestingly, as shown in Ref. 101, the modular S-matrix can be interpreted as a generalization of the quantum Fourier transform based on the representation theory of the quantum double. In fact, this is exactly the transformation that is implementing the chargeon–fluxon duality discussed in Sec. III B, where the chargeons live on the direct lattice and the fluxons live on the reciprocal lattice. To illustrate this explicitly, all we have to show is that the set of characters corresponding to the chargeon transform, under the action of S, into the set of characters corresponding to the fluxon, since these sets form an orthonormal basis for their respective Hilbert space. We first need the representations labeling the two anyons which can be worked out by virtue of Eq. (29), which defines the quantum double action. Using to denote the representation corresponding to the chargeon e and to denote that of the fluxon m, then the image of their 1D quantum double representations, given the action defined in Eq. (29), is given by Table II.
Since these representations are one-dimensional, they yield the characters directly since the characters are computed as the trace of the elements in the representation image. In light of this, we may denote by the set of characters for the chargeon and by the set of characters for the fluxon. It can then be shown that indeed , which proves that S is implementing a duality between the two anyons. Moreover, if we work out the characters corresponding to the dyon ε, we have that , and if we act with S we find that the state is left invariant, i.e., , as expected since ε is a fluxon–chargeon composite so it must be left untouched if we swap the constituent anyons . This duality is exactly the same as the one discussed in Sec. III B 1 for N = 2. Here, we took a different route via the representation theory and the S-matrix and arrived at the same result, without any direct knowledge about the structure of the Hamiltonian, other than its symmetries. Next, we shall consider spin-2 systems which exhibit phases with non-abelian symmetry groups that may be capable of topological quantum computation.
F. Anyons in spin-2 BECs
Unlike the spin-0 and spin-1 condensates, the spin-2, and in fact any spin , exhibit phases whose order parameters are invariant under non-abelian groups. The order parameters can be worked out from the mean-field theory described in Appendix E. In the unbroken phase, such order parameters have rotational symmetry as well as phase invariance , which results in the full symmetry , but if we consider instead the simply connected special unitary representation , the group structure is given by . If this symmetry is broken to a subgroup , the full order parameter manifold is given by the coset space formed by taking the quotient G/H, i.e., . Here, we will direct our attention toward two particular ground state phases, the binary tetrahedral phase and the biaxial-nematic phase. The particle content of the emerging anyon models can be derived by extending the analysis carried out in the abelian case, but as we shall see, these structures are much more complex which gives rise to vast particle spectra. The topological spins of the particles are, again, derived from the character theory in the same way as in Sec. III E.
1. Binary tetrahedral phase and its quantum double
Instead of writing all of these matrices out explicitly as we did in the abelian C2 case, we are simply providing the corresponding character tables as all essential information can be extracted from there. The characters of the one-dimensional UIRs are found trivially since the traces of them are the same as the group elements themselves. As for the two-dimensional representations, we pick one element from each conjugacy class in Eq. (41) which we multiply by each of the one-dimensional UIRs and then take the trace, since the trace is invariant under conjugacy due to its invariance under cyclic permutations. Finally, the characters for the three-dimensional UIR can be found simply by defining a map f that sends each Pauli matrix σi in Eq. (41) to the corresponding three-dimensional Lorentz rotation matrix Ri, since the Lorentz rotations and adhere to the same algebra. The characters for all of these seven representations can be found in Table III. Again, the topological spins of the anyons in the spectrum can be deduced from the character table via the relation , where s is the spin, χ is the character value and d is the dimension. Reading off the first and the second column, corresponding to fluxons and , we note that all characters are integers which means that they are all bosons. Next, we turn to the conjugacy class which has as centralizer group. This can be concluded by noting that all elements in this set are pure rotations around the x, y, and z axes, respectively. Hence, the only elements commuting with a representative from this conjugacy class is a rotation about the same axis. Considering the group structure, there can only be 4 of those (generated by the element itself), which leads us to the conclusion that the centralizer is . Moreover, each element in is its own conjugacy class since it is an abelian group, and consequently, as per Burnside's theorem, there must be four UIRs ( ), which can be found simply by permuting the trivial representation which send all elements to 1, by rotations of . We, thus, obtain the character table presented in Table IV.
. | . | . | . | . | . | . | . |
---|---|---|---|---|---|---|---|
1 | 1 | 1 | 1 | 1 | 1 | 1 | |
1 | 1 | 1 | ω | ω | |||
1 | 1 | 1 | ω | ω | |||
2 | −2 | 0 | −1 | 1 | 1 | −1 | |
2 | −2 | 0 | ω | ||||
2 | −2 | 0 | ω | ||||
3 | 3 | −1 | 0 | 0 | 0 | 0 |
. | . | . | . | . | . | . | . |
---|---|---|---|---|---|---|---|
1 | 1 | 1 | 1 | 1 | 1 | 1 | |
1 | 1 | 1 | ω | ω | |||
1 | 1 | 1 | ω | ω | |||
2 | −2 | 0 | −1 | 1 | 1 | −1 | |
2 | −2 | 0 | ω | ||||
2 | −2 | 0 | ω | ||||
3 | 3 | −1 | 0 | 0 | 0 | 0 |
. | 1 . | x . | x2 . | x3 . |
---|---|---|---|---|
1 | 1 | 1 | 1 | |
1 | i | −1 | −i | |
1 | −1 | 1 | −1 | |
1 | −i | −1 | i |
. | 1 . | x . | x2 . | x3 . |
---|---|---|---|---|
1 | 1 | 1 | 1 | |
1 | i | −1 | −i | |
1 | −1 | 1 | −1 | |
1 | −i | −1 | i |
The self-statistics, and hence the spin, can be read off from the second column in Table IV which reveals that the dyon is a boson, the is a spin- particle, is a fermion and the must be a spin- particle. Finally, by the same argument we applied to deduce the centralizer, we can establish that the centralizer of the remaining conjugacy classes , and must be the six-cycle since and (where ) have order 6. Again, owing to the abelian structure of , it must have six UIRs whose characters are provided in Table V.
. | 1 . | y . | y2 . | y3 . | y4 . | y5 . |
---|---|---|---|---|---|---|
1 | 1 | 1 | 1 | 1 | 1 | |
1 | −1 | |||||
1 | 1 | |||||
1 | −1 | 1 | −1 | 1 | −1 | |
1 | 1 | |||||
1 | −1 |
. | 1 . | y . | y2 . | y3 . | y4 . | y5 . |
---|---|---|---|---|---|---|
1 | 1 | 1 | 1 | 1 | 1 | |
1 | −1 | |||||
1 | 1 | |||||
1 | −1 | 1 | −1 | 1 | −1 | |
1 | 1 | |||||
1 | −1 |
Even here we find some interesting dyonic particles with fractional spin , which can be deduced from the second column. We have now worked out the entire particle spectrum of the quantum double of which is comprised of one vacuum, six pure fluxons, six pure chargeons, and 29 dyons, thus amounting to a total of 42 distinguishable particles. The fusion rules of these anyons can be obtained by first computing the S-matrix according to Eq. (33) and then employing, e.g., the Verlinde equation in Eq. (32). However, computing all such combinations would be a monstrous task so we will not present these here. The underlying principle is the same though as in the much simpler anyon model. Interestingly, as already pointed out in Sec. III C, fusing a fluxon with its anti-partner may result in a particle with Cheshire charge. This possibility stems from the reducibility of the tensored representation space of the two fluxons. If the resulting space has invariant subspaces, these subspaces correspond to multiplets, each of which is labeled by an invariant charge quantum number.
2. Biaxial nematic phase and its quantum double
The conjugacy class structure and the centralizer UIRs of this group can be found by applying the same set of arguments as in the case, so we are simply jumping straight to the character tables here. Again, the abelian fluxons and have the entire group as centralizer whose character table is provided in Table VI. The dyons with flux corresponding to and are consequently all bosons since all of the characters in the first and second column are integers. The centralizer of and is given by , which coincides with the centralizer of , whose character table is already provided in Table IV. Finally, the centralizer for the remaining conjugacy classes , and are given by the cyclic abelian group whose character table is provided in Table VII.
. | . | . | . | . | . | . | . |
---|---|---|---|---|---|---|---|
1 | 1 | 1 | 1 | 1 | 1 | 1 | |
1 | 1 | −1 | 1 | 1 | 1 | −1 | |
1 | 1 | 1 | 1 | −1 | −1 | −1 | |
1 | 1 | −1 | 1 | −1 | −1 | 1 | |
2 | −2 | 0 | 0 | −2 | 0 | 0 | |
4 | −4 | 0 | 0 | 0 | |||
4 | −4 | 0 | 0 | 0 |
. | . | . | . | . | . | . | . |
---|---|---|---|---|---|---|---|
1 | 1 | 1 | 1 | 1 | 1 | 1 | |
1 | 1 | −1 | 1 | 1 | 1 | −1 | |
1 | 1 | 1 | 1 | −1 | −1 | −1 | |
1 | 1 | −1 | 1 | −1 | −1 | 1 | |
2 | −2 | 0 | 0 | −2 | 0 | 0 | |
4 | −4 | 0 | 0 | 0 | |||
4 | −4 | 0 | 0 | 0 |
. | 1 . | z . | z2 . | z3 . | z4 . | z5 . | z6 . | z7 . |
---|---|---|---|---|---|---|---|---|
1 | 1 | 1 | 1 | 1 | 1 | 1 | 1 | |
1 | −1 | 1 | −1 | 1 | −1 | 1 | −1 | |
1 | i | −1 | −i | 1 | i | −1 | −i | |
1 | −i | −1 | i | 1 | −i | −1 | i | |
1 | θ | |||||||
1 | θ | |||||||
1 | θ | |||||||
1 | θ |
. | 1 . | z . | z2 . | z3 . | z4 . | z5 . | z6 . | z7 . |
---|---|---|---|---|---|---|---|---|
1 | 1 | 1 | 1 | 1 | 1 | 1 | 1 | |
1 | −1 | 1 | −1 | 1 | −1 | 1 | −1 | |
1 | i | −1 | −i | 1 | i | −1 | −i | |
1 | −i | −1 | i | 1 | −i | −1 | i | |
1 | θ | |||||||
1 | θ | |||||||
1 | θ | |||||||
1 | θ |
We deduce from these characters non-abelian dyons with fractional topological spins .
3. A note on the quaternionic phase and its quantum double
In a separate work,104 devoted entirely to the model, we analyzed the feasibility of employing the quaternionic phase of a spin-2 BEC as a TQC platform. This phase can be obtained from the one by reducing its rotational symmetry to a four-cycle. For the sake of completeness, we are here outlining some key aspect of this model. One of the primary challenges of developing a theoretical model of a TQC based on the non-abelian phases considered in this work, is that the dimensionality of the fusion spaces are generally larger than two. This is problematic since we only need a two-level system to form a qubit, meaning that the redundant particles in the fusion outcomes will generally absorb amplitude and consequently cause information leakage.82,105 It is, therefore, of interest to find a low-temperature symmetry group whose quantum double contains fusion products that form two-level states. The quaternionic phase offers such a model if one consider a closed subset of the particles in the theory.
A natural question that may arise is why we do not want to utilize all of the particles in a fusion outcome to represent a generic qudit. The issue with this idea, as thoroughly discussed in Ref. 104, is that one would need to have access to a group whose algebra is of a higher dimension. For instance, the braid group of a four-level qudit still only have two generators σ1 and σ2, which is insufficient to span an Bloch sphere. Instead, by considering two qubits, which combined form a four-level system, the braid group has five generators, which may, thus, make universality more attainable. For instance, as already touched upon in Subsection III D and demonstrated in Ref. 100, a topological gravity may be simulated via the quantum pressure in the fluid, which results in additional topologically protected phase rotations which one may supplement any non-universal braid set with. Only one such additional phase gate would be required in order to make the universal. Another advantage of this model is that all fusion rules are multiplicity free. As already argued in Ref. 106, the presence of multiplicities is complicating the calculation of the Clebsch–Gordan coefficients of the theory. This is also apparent in our derivation presented in Appendix D since the number of terms on the right hand side in Eq. (D10) depends on the multiplicity. Only in the multiplicity-free case is an explicit solution accessible.
G. Symmetry-based normal mode analysis
As indicated by the Fourier duality connecting the fluxons and the chargeons, if one of the two is considered to be localized in real space, its dual must be localized in the reciprocal space, with respect to a generalized quantum Fourier transform (the S-matrix). While it is, as per homotopy theory, well established that the fluxons must map onto quantized vortices, what the physical incarnations of chargeons are is less clear. The purpose of this section is, therefore, to offer an interpretation of the chargeon degrees of freedom. We find that, by means of a symmetry analysis, these normal modes correspond to spin-rotations and spin-waves or magnons. These systems also have regular phonon excitations but those are naturally associated with the breaking of translational symmetry. In general, a system in equilibrium exhibits some sort of symmetry that, if broken, excites modes in the system. These solutions transform according to the various UIRs of the, possibly broken, symmetry group, and as such, furnish a basis for the vector spaces the matrices in images of the UIRs act on. Guided by these considerations, we may now proceed and work out the normal modes in a spinor BEC.
1. Tetrahedral phase—An example
What we learn from this is that there is one unique eigenvalue attached to each UIR, and that this eigenvalue must be the same for and . Moreover, the dimension of the subspace Vi, and hence the degeneracy of , must be equal to the dimension of the UIR . We have now reduced the problem of finding the normal modes of the system to finding the various irreducible representations of the underlying symmetry group. Remarkably, the number of distinct eigenvalues and their degeneracies can be completely deduced from the representation theory, without any consideration of the physical parameters; the entire measurable spectrum of possible eigenvalues is classified by the symmetry of the system. Note that this is in complete agreement with the quantum double particle labeling. The fluxons are point particles labeled by the conjugacy classes of the group, and the chargeons are labeled by the various UIRs (or the reverse if we act with the S-matrix and consider the system in reciprocal space).
IV. VORTEX ANYONS IN SUPERFLUID GASES OF COLD ATOMS
To draw a connection between the theoretical concepts and experiments, we briefly mention two closely related physical systems, superfluid Fermi gases and Bose–Einstein condensates, which may be able to host non-abelian anyons inside the cores of quantized vortices. We will first discuss abelian vortices, routinely created and observed in experiments, and how to manipulate them before contemplating their respective non-abelian extensions yet to be realized in the laboratory.
A. Quantized vortices in cold atomic superfluids
Scalar vortices were created and observed in Bose–Einstein condensates in 1999107 and in superfluid Fermi gas in 2005.108 Since then, they have been routinely observed in numerous cold atom laboratories using a variety of techniques. Further details about vortex experiments in cold atom superfluids may be found for instance in Refs. 109–111.
In both bosonic and fermionic cases, the vortices, if left to equilbrate, will typically arrange into a regular Wigner-crystalline triangular vortex lattice due to the repulsive 2D-Coulomb-like interaction between the vortices.
B. Non-abelian vortices in a superfluid Fermi gas
Here, we briefly outline how the non-abelian anyons may emerge in the vortex cores of chiral p-wave paired superfluid Fermi gas. This situation is generic and believed to be relevant to other topological fermionic superfluids such as in liquid 3He and certain Type II superconductors.13,23,26,27,112–114
From theoretical perspective, vortices must be well separated to avoid quasi-particle tunneling between vortex cores, which would cause energy splitting of the Majorana pairs. From an experimental perspective, the greatest unresolved challenge is to realize a suitable topological superfluid such as a spinless p-wave paired superfluid phase of a Fermi gas. Ultracold Fermi gases near a p-wave Feshbach resonance, have been studied experimentally116–123 and offer a potential route to realizing chiral p-wave superfluidity, however, they suffer from inelastic losses which are enhanced near the resonance.124 Lower-dimensional gases,125–129 optical lattices130 and even nonequilibrium131 systems provide promising opportunities for studying Fermi gases with resonantly enhanced p-wave interactions and remain the subject of ongoing research.
C. Non-abelian vortices in a superfluid Bose gas
In contrast to the fermionic case where the role of the Majorana vortices is merely to trap and host the core localized quasi-particles, which define the qubit of the non-abelian fusion algebra, in the quantum double construction of spinor BECs, both the carrier vortices (fluxons) and the core localized quasi-particles (chargeons) participate in the fusion process. As such, the Majorana zero modes may be viewed as the counterparts of pure chargeons of the quantum double.
D. Experiment considerations regarding creation, fusion, and braiding
It is useful to first consider the most abundant laboratory system of a simple scalar BEC with a quantized abelian vortex fluxon. Such vortices are readily produced in research laboratories via a variety of methods,109–111 including the so-called “chopsticks” method,135 which enables deterministic production of vortices and control over their positions, thereby conceptually facilitating pair creation, braiding, and fusion protocols for vortex anyons. Recently, this method has been utilized to develop a programable vortex collider in a superfluid Fermi gas.136 In principle, these techniques can be extended to larger numbers of vortices using optical tweezer arrays, as being deployed for storing Rydberg atom qubits in neutral atom quantum computers.137–139
As the vortices are moved around during braiding, one potential concern is the effect of phonon excitations. Although phonons cannot change the topology of isolated vortices, it is in principle possible for a high-energy phonon to split into a vortex–antivortex pair via a Sauter–Schwinger-like pair creation process and such vortices could then accidentally become braided as the computation is carried out leading to quasi-particle poisoning of the computation. The probability of such processes should, however, be negligible at low enough temperatures achievable in the experiments. In the spinor BECs, it is also conceivable that an absorption of a phonon by the vortex might affect the atomic probability distribution in the different hyperfine spin levels of the core localized modes. This could potentially alter the amplitudes of the quantum double fusion rules. However, we are not aware of any works that would have addressed this issue.
Regarding the readout of the fusion outcome, the greatest experimental challenge is to observe the vortex core localized quasi-particles, which has not been achieved to date even in the case of a scalar BEC. It was only recently discovered140 that even in a scalar BEC a vortex core hosts a zero energy BdG quasi-particle mode known as a kelvon.141,142 Such a kelvon can be regarded as a quantum mechanical equivalent of a classical Kelvin wave.143 In a 3D scalar BEC, kelvons have been indirectly observed144 as they distort the shape of a vortex line into a helix. However, in 2D, it is not clear how to experimentally resolve the zero energy kelvon, which can be viewed as a quantum depleted non-condensate atom that is trapped by the vortex core and as such is inseparable from the vortex fluxon. It may, thus, be tempting to interpret the kelvon as the charge of a quantum double anyon. Indeed, from our analysis we found that all pure fluxons are in fact bosons, while the dyons generally have fractional charge. Therefore, the non-abelian anyons in a spinor BEC are always charge–flux composites. We may, therefore, conjecture that the chargeon part of a dyon corresponds to the spin mode of the kelvon attached to the vortex. This view is supported by numerical experiments carried out in Ref. 57, where spin waves were observed after a fusion event involving two vortices. Even if a vortex is annihilated by an anti-vortex, they may still have a charge that survives the fusion process. In a realistic model of TQC based on vortices in a spinor BEC, one would then build the qubits from vortices and their kelvons.
Thus, the technologies to create, braid, and fuse vortex anyons have been demonstrated in the case of abelian vortices. There are two remaining major problems that need to be resolved in order to achieve a complete experimental demonstration of a non-abelian vortex based topological qubit in a cold atom superfluid. The first is the stabilization of a ground state whose order parameter would support non-abelian vortices as its natural excitations. For instance, it is not presently known if the non-abelian cyclic, biaxial nematic, or the quaternion phases can be stabilized as ground states.133 The second, which has already been discussed, is the development of experimental techniques that allow probing the kelvon and to measure its charge, a task that is yet to be realized even for vortices in existing quasi-two-dimensional scalar BECs.
V. DISCUSSION
We have aimed at providing a self-contained account of the emergent low-temperature quantum double structure in spinor BECs, a connection that is not well covered in the literature to date. It is well understood that spinor BECs may support non-abelian anyons in the form of quantized vortices (fluxons). However, as emphasized in this paper, this is just one side of the story. The existence of a particle-vortex duality, whose structure is made exact within the quantum double framework, implies that the description of vortices can be mapped onto a dual side, which describes a different type of particle (chargeons). We have shown that the chargeons belonging to the dual side correspond to spin waves and spin rotation modes. As illustrated, despite being non-abelian, the fluxons only have fractional mutual statistics, whereas their self-statistics is bosonic. The same is true for the chargeons since their algebraic structure is, owing to the duality, equivalent. However, the quantum double structure also accommodate charge–flux composites which, as demonstrated, may indeed have fractional self-statistics.
The strong interest in non-abelian anyons is primarily due to their potential to realize fault-tolerant quantum information processing. Up until now, the main focus has been on fermionic systems, and in particular systems that may host Majorana fermion quasi-particles. Spinor BECs, on the other hand, appear to have been overlooked in this context, which prompted us to carry out this work with the aim of shedding light on their potential for TQC. Theoretically, as we discussed in Sec. III F 3, one of the main hurdles in the development of a spinor BEC-based TQC platform is that the fusion spaces generally are large due to the high dimensionality. To construct a qubit, only two states are needed, which implies that information leakage82,105 may occur into the states not belonging to the computational space. Another issue is the possibility of multiplicities in the fusion rules, which complicates severely the calculation of the Clebsch–Gordan coefficients106 and, thus, the braid group matrices. Similar issues are, of course, also present in the anyon models but mainly for larger values of k. To reduce the chance of having to face such problems, it is useful to consider smaller groups since that will restrict the dimensionality of the Hilbert spaces spanned by the anyons of the theory. For instance, the quaternionic subgroup , which was considered in a separate work,104 can be obtained by breaking further the rotational part of the subgroup to a four-cycle. For this group, it is possible to define multiplicity free qubits that are not accompanied by additional non-computational states that generally cause leakage.
Finally, although a scalable fault tolerant topological quantum computer based on non-abelian vortex anyons remains out of reach, a neutral atom superfluid prepared in a non-abelian ground state could certainly facilitate a beautiful demonstration of the fundamental principles of topological quantum computation utilizing a few topologically protected logical qubits.
ACKNOWLEDGMENTS
This research was supported by the Australian Research Council Future Fellowship FT180100020, and was funded by the Australian Government.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Emil Génetay Johansen: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal). Chris Vale: Supervision (supporting); Validation (equal); Writing – review & editing (equal). Tapio Simula: Conceptualization (equal); Formal analysis (supporting); Funding acquisition (lead); Investigation (supporting); Methodology (equal); Supervision (lead); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article.