A picture of me.

Tom Hodson

Maker, Baker Programmer Reformed Physicist RSE@ECMWF



1 Introduction

Interacting Quantum Many-Body Systems

When you take many objects and let them interact together, complex behaviours can emerge. It is often easier to describe these behaviours in terms of properties of the group rather than properties of the individual objects. A flock of starlings like that of fig. 1 is a good example. If you were to sit and watch a flock like this, you’d see that it has a distinct outline, that waves of density will sometimes propagate through the closely packed birds and that the flock seems to respond to predators as a distinct object. The natural description of this phenomenon is in terms of the flock, not the individual birds.

A flock is an emergent phenomenon. The starlings are only interacting with their immediate six or seven neighbours  [1,2], what a physicist would call a local interaction. There is much philosophical debate about how exactly to define emergence  [3,4] but, for our purposes, it is enough to say that emergence is the fact that the aggregate behaviour of many interacting objects may necessitate a radically different description from that of the individual objects.

Figure 1: A murmuration of starlings. Dorset, UK. Credit Tanya Hart, “Studland Starlings”, 2017, CC BY-SA 3.0

To give an example closer to the topic at hand, our understanding of thermodynamics began with bulk properties like heat, pressure, energy and temperature  [5]. It was only later that we gained an understanding of how these properties emerge from microscopic interactions between very large numbers of particles  [6].

At its heart, condensed matter is the study of the behaviours that can emerge from large numbers of interacting quantum objects at low energy. From these three ingredients,: a large number of objects, those objects being quantum and the presence of interactions between the objects, nature builds all manner of weird and wonderful things, see fig. 2 for examples. When these three properties are all present and important, we call it an interacting quantum many-body system. Such systems will be the focus of this thesis.

Historically, we first made headway by ignoring interactions and quantum properties and looking at purely many-body systems. The ideal gas law and the Drude classical electron gas  [7] are good examples. Including interactions leads to the Ising model  [8], Landau theory  [9] and the classical theory of phase transitions  [10]. In contrast, condensed matter theory got its start in quantum many-body theory where the only electron-electron interaction considered is the Pauli exclusion principle 1. Bloch’s theorem  [11], the core result of band theory, predicted the properties of non-interacting electrons in crystal lattices. In particular, it predicted that band insulators arise when the electrons bands are filled, leaving the fermi level in a bandgap  [7]. In the same vein, advances were made in understanding the quantum origins of magnetism, including ferromagnetism and antiferromagnetism  [12].

The development of Landau-Fermi liquid theory explained why band theory works so well even when an analysis of the relevant energies suggests that it should not  [13]. Landau-Fermi liquid theory demonstrates that, in many cases where electron-electron interactions are significant, the system can still be described in terms of generalised non-interacting quasiparticles. This description is applicable when the properties of the quasiparticles in the interacting system can be smoothly connected to the free fermions of the non-interacting system.

However, there are systems where even Landau-Fermi liquid theory fails. An effective theoretical description of these systems must include electron-electron correlations. They are thus called strongly correlated materials  [14]. The canonical examples are superconductivity  [15], the fractional quantum Hall effect  [16] and the Mott insulators  [17,18]. We’ll start by looking at the latter but shall see that there are many links between the three topics.

Mott Insulators

Mott Insulators (MIs) are remarkable because their electrical insulator properties come not from having filled bands but from electron-electron interactions other than Pauli exclusion. Electrical conductivity, the bulk movement of electrons, requires both that there are electronic states very close in energy to the ground state and that those states are delocalised so that they can contribute to macroscopic transport. Band insulators are systems whose Fermi level falls within a gap in the density of states: they fail the first criteria. Band insulators derive their insulating character from the characteristics of the underlying lattice. Another class of insulator, the Anderson insulators, are disordered so only have localised electronic states near the fermi level. They therefore fail the second criteria. In a later section, I will discuss Anderson insulators and the disorder that drives them. Both band and Anderson insulators occur without electron-electron interactions. MIs, by contrast, require a many-body picture to understand and thus elude band theory and single-particle methods.

Figure 2: Three key adjectives. Many-Body: systems considered in the limit of large numbers of particles. Quantum: objects whose behaviour requires quantum mechanics to describe accurately. Interacting: the constituent particles of the system affect one another via forces, either directly or indirectly. When taken together, these three properties can give rise to strongly correlated materials.

The theory of MIs developed out of the observation that band theory erroneously predicts that many transition metal oxides are conductive  [19]. It was suggested that electron-electron interactions were the cause of this effect  [20]. Interest grew further with the discovery of high temperature superconductivity in the cuprates in 1986  [21] which is believed to arise as the result of a doped MI state  [22].

The canonical toy model of the MI is the Hubbard model  [2325] of spin-\(1/2\) fermions hopping on the lattice with hopping parameter \(t\) and electron-electron repulsion \(U\), it reads

\[ H_{\mathrm{H}} = -t \sum_{\langle i,j \rangle \alpha} c^\dagger_{i\alpha} c^{\phantom{\dagger}}_{j\alpha} + U \sum_i n_{i\uparrow} n_{i\downarrow} - \mu \sum_{i,\alpha} n_{i\alpha},\]

where \(c^\dagger_{i\alpha}\) creates a spin \(\alpha\) electron at site \(i\) and the number operator \(n_{i\alpha}\) measures the number of electrons with spin \(\alpha\) at site \(i\). The sum runs over lattice neighbours \(\langle i,j \rangle\) including both \(\langle i,j \rangle\) and \(\langle j,i \rangle\) so that the model is Hermitian.

In the non-interacting limit \(U \ll t\), the model reduces to free fermions and the many-body ground state is a separable product of Bloch waves filled up to the Fermi level. On the other hand, the ground state in the interacting limit \(U \gg t\) is a direct product of the local Hilbert spaces \(|0\rangle, |\uparrow\rangle, |\downarrow\rangle, |\uparrow\downarrow\rangle\). At half-filling, one electron per site, each site becomes a local moment in the reduced Hilbert space \(|\uparrow\rangle, |\downarrow\rangle\) and thus acts like a spin-\(1/2\)  [26].

The Mott insulating phase occurs at half-filling \(\mu = \tfrac{U}{2}\). Here the model can be rewritten in a symmetric form

\[ H_{\mathrm{H}} = -t \sum_{\langle i,j \rangle \alpha} c^\dagger_{i\alpha} c^{\phantom{\dagger}}_{j\alpha} + U \sum_i (n_{i\uparrow} - \tfrac{1}{2})(n_{i\downarrow} - \tfrac{1}{2}).\]

The basic reason that the half-filled state is insulating seems trivial. Any excitation must include states of double occupancy that cost energy \(U\). Hence, the system has a finite bandgap and is an interaction-driven MI. Depending on the lattice, the local moments may then order antiferromagnetically. Originally it was proposed that this antiferromagnetic (AFM) order was actually the reason for the insulating behaviour. This would make sense since AFM order doubles the unit cell and can turn a system into a band insulator with an even number of electrons per unit cell  [27]. However, MIs have been found without magnetic order  [28,29]. Instead, the local moments may form a highly entangled state known as a Quantum Spin Liquid (QSL), which will be discussed shortly.

Various theoretical treatments of the Hubbard model have been made, including those based on Fermi liquid theory, mean field treatments, the local density approximation  [30], dynamical mean-field theory  [31], density matrix renormalisation group methods  [3234] and Markov chain Monte Carlo  [3537]. None of these approaches are perfect. Strong correlations are poorly described by Landau-Fermi liquid theory and local density approximation approaches while mean field approximations do poorly in low dimensional systems. This theoretical difficulty has made the Hubbard model a target for cold atom simulations  [38].

From here, the discussion will branch in two directions. First, I will discuss a limit of the Hubbard model called the Falicov-Kimball model. Second, I will look at QSLs and the Kitaev honeycomb model.

The Falicov-Kimball Model

Figure 3: The Falicov-Kimball model can be viewed as a model of classical spins S_i coupled to spinless fermions \hat{c}_i where the fermions are mobile with hopping t and the fermions are coupled to the spins by an Ising type interaction with strength U.

The Falicov-Kimball (FK) model was originally introduced to describe the metal-insulator transition in f-electron systems  [25,39]. Shown graphically in fig. 3, the FK model is the limit of the Hubbard model as the mass of one of the spin states of the electron is taken to infinity. This gives a model with two fermion species, one itinerant and one entirely immobile. The number operators for the immobile fermions are therefore conserved quantities and can be treated like classical degrees of freedom. For our purposes, it will be useful to replace the immobile fermions with a classical Ising background field \(S_i = \pm1\). At half filing and with this substitution, the Hamiltonian reads

\[\begin{aligned} H_{\mathrm{FK}} = & -\;t \sum_{\langle i,j \rangle} c^\dagger_{i}c^{\phantom{\dagger}}_{j} + \;U \sum_{i} S_i\;(c^\dagger_{i}c^{\phantom{\dagger}}_{i} - \tfrac{1}{2}). \\ \end{aligned}\]

The physics of states near the metal-insulator transition is still poorly understood  [40,41]. As a result, the FK model provides a rich test bed to explore interaction-driven metal-insulator transition physics. Despite its simplicity, the model has a rich phase diagram in \(D \geq 2\) dimensions. It shows a Mott insulator transition even at high temperature, similar to the corresponding Hubbard model  [42]. In 1D, the ground state phenomenology as a function of filling can be rich  [43], but the system is disordered for all \(T > 0\)  [44]. The model has also been a test-bed for many-body methods. Interest took off when an exact dynamical mean-field theory solution in the infinite dimensional case was found  [4548].

In chapter 3, I will introduce a generalised Falicov-Kimball model in 1D I call the Long-Range Falicov-Kimball model. With the addition of long-range interactions in the background field, the model shows a rich phase diagram like its higher dimensional cousins. Our goal is to understand the Mott transition in more detail, the phase transition into a charge density wave state and how the localisation properties of the fermionic sector behave in 1D. I was particularly interested to see if correlations in the disorder potential are enough to bring about localisation effects, such as mobility edges, that are normally only seen in higher dimensions. I use an exact Markov chain Monte Carlo method to map the phase diagram and compute the energy-resolved localisation properties of the fermions. We observe what appears to be a hint of coexisting localised and delocalised states. However, after careful comparison to an Anderson model of uncorrelated binary disorder about a background charge density wave field, we confirm that the fermionic sector does fully localise at larger system sizes as expected for 1D systems.

Quantum Spin Liquids

Turning to the other key topic of this thesis, we have already discussed the AFM ordering of local moments in the Mott insulating state. Landau-Ginzburg-Wilson theory characterises phases of matter as inextricably linked to the emergence of long-range order via a spontaneously broken symmetry. Within this paradigm, we would not expect any interesting phases of matter not associated with AFM or other long-range order. However, Anderson first proposed in 1973  [49] that, if long-range order is suppressed by some mechanism, it might lead to a liquid-like state even at zero temperature: a QSL.

This QSL state would exist at zero or very low temperatures. Therefore, we would expect quantum effects to be very strong, which will have far reaching consequences. It was the discovery of a different phase, however, that really kickstarted interest in the topic. The fractional quantum Hall state, discovered in the 1980s  [50] is an explicit example of an interacting electron system that falls outside of the Landau-Ginzburg-Wilson paradigm2. It shares many phenomenological properties with the QSL state. They both exhibit fractionalised excitations, braiding statistics and non-trivial topological properties  [55]. The many-body ground state of such systems acts as a complex and highly entangled vacuum. This vacuum can support quasiparticle excitations with properties unbound from that of the Dirac fermions of the standard model.

How do we actually make a QSL? Frustration is one mechanism that we can use to suppress magnetic order in spin models  [56]. Frustration can be geometric. Triangular lattices, for instance, cannot support AFM order. It can also come about as a result of spin-orbit coupling or other physics. There are also other routes to QSLs besides frustrated spin systems that we will not discuss here  [5759].

Figure 4: How Kitaev materials fit into the picture of strongly correlated systems. Interactions are required to open a Mott gap and localise the electrons into local moments, while spin-orbit correlations are required to produce the strongly anisotropic spin-spin couplings of the Kitaev model. Reproduced from  [56].

Spin-orbit coupling is a relativistic effect that, very roughly, corresponds to the fact that in the frame of reference of a moving electron the electric field of nearby nuclei looks like a magnetic field to which the electron spin couples. This couples the spatial and spin parts of the electron wavefunction. The lattice structure can therefore influence the form of the spin-spin interactions, leading to spatial anisotropy in the effective interactions. This spatial anisotropy can frustrate an MI state  [60,61] leading to more exotic ground states than the AFM order we have seen so far. As with the Hubbard model, interaction effects are only strong or weak in comparison to the bandwidth or hopping integral \(t\). Hence, we will see strong frustration in materials with strong spin-orbit coupling \(\lambda\) relative to their bandwidth \(t\).

In certain transition metal based compounds, such as those based on iridium and ruthenium, the lattice structure, strong spin-orbit coupling and narrow bandwidths lead to effective spin-\(\tfrac{1}{2}\) Mott insulating states with strongly anisotropic spin-spin couplings. These transition metal compounds, known as Kitaev materials, draw their name from the celebrated Kitaev Honeycomb (KH) model which is expected to model their low temperature behaviour  [56,6265].

At this point, we can sketch out a phase diagram like that of fig. 4. When both electron-electron interactions \(U\) and spin-orbit couplings \(\lambda\) are small relative to the bandwidth \(t\), we recover standard band theory of band insulators and metals. In the upper left, we have the simple Mott insulating state as described by the Hubbard model. In the lower right, strong spin-orbit coupling gives rise to topological insulators characterised by symmetry protected edge modes and non-zero Chern number. Kitaev materials occur in the region where strong electron-electron interaction and spin-orbit coupling interact. See ref.  [66] for a more expansive version of this diagram.

The KH model  [67] was one of the early exactly solvable spin models with a QSL ground state. It is defined on the 2D honeycomb lattice and provides an exactly solvable model that can be reduced to a free fermion problem via a mapping to Majorana fermions. This yields an extensive number of static \(\mathbb Z_2\) fluxes tied to an emergent gauge field. The model is remarkable not only for its QSL ground state, but also for its fractionalised excitations with non-trivial braiding statistics. It has a rich phase diagram hosting gapless, Abelian and non-Abelian phases  [68] and a finite temperature phase transition to a thermal metal state  [69]. It has been proposed that its non-Abelian excitations could be used to support robust topological quantum computing  [7072].

The KH and FK models have quite a bit of conceptual overlap. They can both be seen as models of spinless fermions coupled to a classical Ising background field. This is what makes them exactly solvable. At finite temperatures, fluctuations in their background fields provide an effective disorder potential for the fermionic sector, so both models can be studied at finite temperature with Markov chain Monte Carlo methods  [69,73].

As Kitaev points out in his original paper, the KH model remains solvable on any trivalent \(z=3\) graph which can be three-edge-coloured. Indeed, many generalisations of the model exist  [7478]. Notably, the Yao-Kivelson model  [79] introduces triangular plaquettes to the honeycomb lattice leading to spontaneous chiral symmetry breaking. These extensions all retain translation symmetry. This is likely because edge-colouring, finding the ground state and understanding the QSL properties are much harder without it  [80,81]. Undeterred, this gap led us to wonder what might happen if we remove translation symmetry from the Kitaev model. This would be a model of a trivalent, highly bond anisotropic but otherwise amorphous material.

Amorphous materials do not have long-range lattice regularities but may have short-range regularities in the lattice structure, such as fixed coordination number \(z\) as in some covalent compounds. The best examples are amorphous silicon and germanium with \(z=4\) which are used to make thin-film solar cells  [82,83]. Recently, it has been shown that topological insulating (TI) phases can exist in amorphous systems. Amorphous TIs are characterised by similar protected edge states to their translation invariant cousins and generalised topological bulk invariants  [8490]. However, research on amorphous electronic systems has mostly focused on non-interacting systems with a few exceptions, for example, to account for the observation of superconductivity  [9195] in amorphous materials or very recently to understand the effect of strong electron repulsion in TIs  [96].

Amorphous magnetic systems have been investigated since the 1960s, mostly through the adaptation of theoretical tools developed for disordered systems  [97100] and with numerical methods  [101,102]. Research on classical Heisenberg and Ising models accounts for the observed behaviour of ferromagnetism, disordered antiferromagnetism and widely observed spin glass behaviour  [103]. However, the role of spin-anisotropic interactions and quantum effects in amorphous magnets has not been addressed.

In chapter 4, I will address the question of whether frustrated magnetic interactions on amorphous lattices can give rise to genuine quantum phases, i.e., to long-range entangled QSL  [49,104106]. We will find that the answer is yes. I will introduce the amorphous Kitaev model, a generalisation of the KH model to random lattices with fixed coordination number three. I will show that this model is a solvable, amorphous, chiral spin liquid. As with the Yao-Kivelson model  [79], the amorphous Kitaev model retains its exact solubility but the presence of plaquettes with an odd number of sides leads to a spontaneous breaking of time reversal symmetry. I will confirm prior observations that the form of the ground state is relatively simple  [77,107] and unearth a rich phase diagram displaying Abelian as well as a non-Abelian chiral spin liquid phases. Furthermore, I will show that the system undergoes a finite-temperature phase transition to a thermal metal state and discuss possible experimental realisations.

The next chapter, Chapter 2, will introduce some necessary background to the FK model, the KH model, and disorder and localisation.

Next Chapter: 2 Background

Bibliography

[1]
A. J. King and D. J. Sumpter, Murmurations, Current Biology 22, R112 (2012).
[2]
M. Ballerini et al., Interaction Ruling Animal Collective Behavior Depends on Topological Rather Than Metric Distance: Evidence from a Field Study, Proceedings of the National Academy of Sciences 105, 1232 (2008).
[3]
P. W. Anderson, More Is Different, Science 177, 393 (1972).
[4]
S. Kivelson and S. A. Kivelson, Defining Emergence in Physics, Npj Quant Mater 1, 1 (2016).
[5]
W. M. Saslow, A History of Thermodynamics: The Missing Manual, Entropy (Basel) 22, 77 (2020).
[6]
D. Flamm, History and Outlook of Statistical Physics, arXiv:physics/9803005.
[7]
N. W. Ashcroft and N. D. Mermin, Solid State Physics (Holt, Rinehart and Winston, 1976).
[8]
E. Ising, Beitrag zur Theorie des Ferromagnetismus, Z. Physik 31, 253 (1925).
[9]
L. D. Landau, On the Theory of Phase Transitions. I., Phys. Z. Sowjet. 11, 19 (1937).
[10]
[11]
[12]
S. Blundell, Magnetism in Condensed Matter (OUP Oxford, 2001).
[13]
[14]
E. Morosan, D. Natelson, A. H. Nevidomskyy, and Q. Si, Strongly Correlated Materials, Adv. Mater. 24, 4896 (2012).
[15]
J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Microscopic Theory of Superconductivity, Phys. Rev. 106, 162 (1957).
[16]
D. E. Feldman and B. I. Halperin, Fractional Charge and Fractional Statistics in the Quantum Hall Effects, Rep. Prog. Phys. 84, 076501 (2021).
[17]
[18]
M. P. A. Fisher, Mott Insulators, Spin Liquids and Quantum Disordered Superconductivity, in Aspects Topologiques de La Physique En Basse Dimension. Topological Aspects of Low Dimensional Systems, edited by A. Comtet, T. Jolicœur, S. Ouvry, and F. David, Vol. 69 (Springer Berlin Heidelberg, Berlin, Heidelberg, 1999), pp. 575–641.
[19]
J. H. de Boer and E. J. W. Verwey, Semi-Conductors with Partially and with Completely Filled 3d-Lattice Bands, Proc. Phys. Soc. 49, 59 (1937).
[20]
N. F. Mott and R. Peierls, Discussion of the Paper by de Boer and Verwey, Proc. Phys. Soc. 49, 72 (1937).
[21]
J. G. Bednorz and K. A. Müller, Possible highTc Superconductivity in the Ba−La−Cu−O System, Z. Physik B - Condensed Matter 64, 189 (1986).
[22]
P. A. Lee, N. Nagaosa, and X.-G. Wen, Doping a Mott Insulator: Physics of High-Temperature Superconductivity, Rev. Mod. Phys. 78, 17 (2006).
[23]
M. C. Gutzwiller, Effect of Correlation on the Ferromagnetism of Transition Metals, Phys. Rev. Lett. 10, 159 (1963).
[24]
J. Kanamori, Electron Correlation and Ferromagnetism of Transition Metals, Progress of Theoretical Physics 30, 275 (1963).
[25]
Hubbard, J., Electron Correlations in Narrow Energy Bands, Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences 276, 238 (1963).
[26]
J. Hubbard, Electron Correlations in Narrow Energy Bands III. An Improved Solution, Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences 281, 401 (1964).
[27]
N. Mott, Metal-Insulator Transitions (CRC Press, London, 1990).
[28]
K. T. Law and P. A. Lee, 1t-TaS2 as a Quantum Spin Liquid, Proceedings of the National Academy of Sciences 114, 6996 (2017).
[29]
A. Ribak, I. Silber, C. Baines, K. Chashka, Z. Salman, Y. Dagan, and A. Kanigel, Gapless Excitations in the Ground State of 1t-TaS2, Phys. Rev. B 96, 195131 (2017).
[30]
J. C. Slater, Magnetic Effects and the Hartree-Fock Equation, Phys. Rev. 82, 538 (1951).
[31]
M. Greiner, O. Mandel, T. Esslinger, T. W. Hänsch, and I. Bloch, Quantum Phase Transition from a Superfluid to a Mott Insulator in a Gas of Ultracold Atoms, Nature 415, 39 (2002).
[32]
K. A. Hallberg, New Trends in Density Matrix Renormalization, Advances in Physics 55, 477 (2006).
[33]
U. Schollwöck, The Density-Matrix Renormalization Group, Rev. Mod. Phys. 77, 259 (2005).
[34]
S. R. White, Density Matrix Formulation for Quantum Renormalization Groups, Phys. Rev. Lett. 69, 2863 (1992).
[35]
R. Blankenbecler, D. J. Scalapino, and R. L. Sugar, Monte Carlo Calculations of Coupled Boson-Fermion Systems. I, Phys. Rev. D 24, 2278 (1981).
[36]
[37]
S. R. White, D. J. Scalapino, R. L. Sugar, E. Y. Loh, J. E. Gubernatis, and R. T. Scalettar, Numerical Study of the Two-Dimensional Hubbard Model, Phys. Rev. B 40, 506 (1989).
[38]
A. Mazurenko, C. S. Chiu, G. Ji, M. F. Parsons, M. Kanász-Nagy, R. Schmidt, F. Grusdt, E. Demler, D. Greif, and M. Greiner, A Cold-Atom Fermi–Hubbard Antiferromagnet, Nature 545, 462 (2017).
[39]
L. M. Falicov and J. C. Kimball, Simple Model for Semiconductor-Metal Transitions: SmB6 and Transition-Metal Oxides, Phys. Rev. Lett. 22, 997 (1969).
[40]
D. Belitz and T. R. Kirkpatrick, The Anderson-Mott Transition, Rev. Mod. Phys. 66, 261 (1994).
[41]
D. M. Basko, I. L. Aleiner, and B. L. Altshuler, Metal–Insulator Transition in a Weakly Interacting Many-Electron System with Localized Single-Particle States, Annals of Physics 321, 1126 (2006).
[42]
U. Brandt and C. Mielsch, Thermodynamics and Correlation Functions of the Falicov-Kimball Model in Large Dimensions, Z. Physik B - Condensed Matter 75, 365 (1989).
[43]
C. Gruber, J. Iwanski, J. Jedrzejewski, and P. Lemberger, Ground States of the Spinless Falicov-Kimball Model, Phys. Rev. B 41, 2198 (1990).
[44]
T. Kennedy and E. H. Lieb, An Itinerant Electron Model with Crystalline or Magnetic Long Range Order, Physica A: Statistical Mechanics and Its Applications 138, 320 (1986).
[45]
A. E. Antipov, E. Gull, and S. Kirchner, Critical Exponents of Strongly Correlated Fermion Systems from Diagrammatic Multiscale Methods, Phys. Rev. Lett. 112, 226401 (2014).
[46]
T. Ribic, G. Rohringer, and K. Held, Nonlocal Correlations and Spectral Properties of the Falicov-Kimball Model, Phys. Rev. B 93, 195105 (2016).
[47]
J. K. Freericks and V. Zlatić, Exact Dynamical Mean-Field Theory of the Falicov-Kimball Model, Rev. Mod. Phys. 75, 1333 (2003).
[48]
A. J. Herrmann, N. Tsuji, M. Eckstein, and P. Werner, Nonequilibrium Dynamical Cluster Approximation Study of the Falicov-Kimball Model, Phys. Rev. B 94, 245114 (2016).
[49]
P. W. Anderson, Resonating Valence Bonds: A New Kind of Insulator?, Materials Research Bulletin 8, 153 (1973).
[50]
[51]
S. M. Girvin and A. H. MacDonald, Off-Diagonal Long-Range Order, Oblique Confinement, and the Fractional Quantum Hall Effect, Phys. Rev. Lett. 58, 1252 (1987).
[52]
[53]
[54]
X.-G. Wen, Quantum Orders and Symmetric Spin Liquids, Phys. Rev. B 65, 165113 (2002).
[55]
C. Broholm, R. J. Cava, S. A. Kivelson, D. G. Nocera, M. R. Norman, and T. Senthil, Quantum Spin Liquids, Science 367, eaay0668 (2020).
[56]
S. Trebst and C. Hickey, Kitaev Materials, Physics Reports 950, 1 (2022).
[57]
L. Balents, M. P. A. Fisher, and C. Nayak, Nodal Liquid Theory of the Pseudo-Gap Phase of High-Tc Superconductors, Int. J. Mod. Phys. B 12, 1033 (1998).
[58]
L. Balents, M. P. A. Fisher, and C. Nayak, Dual Order Parameter for the Nodal Liquid, Phys. Rev. B 60, 1654 (1999).
[59]
H.-H. Lin, L. Balents, and M. P. A. Fisher, Exact SO(8) Symmetry in the Weakly-Interacting Two-Leg Ladder, Phys. Rev. B 58, 1794 (1998).
[60]
[61]
G. Khaliullin, Orbital Order and Fluctuations in Mott Insulators, Progress of Theoretical Physics Supplement 160, 155 (2005).
[62]
G. Jackeli and G. Khaliullin, Mott Insulators in the Strong Spin-Orbit Coupling Limit: From Heisenberg to a Quantum Compass and Kitaev Models, Physical Review Letters 102, 017205 (2009).
[63]
M. Hermanns, I. Kimchi, and J. Knolle, Physics of the Kitaev Model: Fractionalization, Dynamic Correlations, and Material Connections, Annual Review of Condensed Matter Physics 9, 17 (2018).
[64]
S. M. Winter, A. A. Tsirlin, M. Daghofer, J. van den Brink, Y. Singh, P. Gegenwart, and R. Valentí, Models and Materials for Generalized Kitaev Magnetism, J. Phys.: Condens. Matter 29, 493002 (2017).
[65]
H. Takagi, T. Takayama, G. Jackeli, G. Khaliullin, and S. E. Nagler, Concept and Realization of Kitaev Quantum Spin Liquids, Nat Rev Phys 1, 4 (2019).
[66]
W. Witczak-Krempa, G. Chen, Y. B. Kim, and L. Balents, Correlated Quantum Phenomena in the Strong Spin-Orbit Regime, Annual Review of Condensed Matter Physics 5, 57 (2014).
[67]
A. Kitaev, Anyons in an Exactly Solved Model and Beyond, Annals of Physics 321, 2 (2006).
[68]
J. Knolle, D. L. Kovrizhin, J. T. Chalker, and R. Moessner, Dynamics of Fractionalization in Quantum Spin Liquids, Phys. Rev. B 92, 115127 (2015).
[69]
C. N. Self, J. Knolle, S. Iblisdir, and J. K. Pachos, Thermally Induced Metallic Phase in a Gapped Quantum Spin Liquid - a Monte Carlo Study of the Kitaev Model with Parity Projection, Phys. Rev. B 99, 045142 (2019).
[70]
A. Yu. Kitaev, Fault-Tolerant Quantum Computation by Anyons, Annals of Physics 303, 2 (2003).
[71]
M. Freedman, A. Kitaev, M. Larsen, and Z. Wang, Topological Quantum Computation, Bull. Amer. Math. Soc. 40, 31 (2003).
[72]
C. Nayak, S. H. Simon, A. Stern, M. Freedman, and S. Das Sarma, Non-Abelian Anyons and Topological Quantum Computation, Rev. Mod. Phys. 80, 1083 (2008).
[73]
A. E. Antipov, Y. Javanmard, P. Ribeiro, and S. Kirchner, Interaction-Tuned Anderson Versus Mott Localization, Phys. Rev. Lett. 117, 146601 (2016).
[74]
G. Baskaran, S. Mandal, and R. Shankar, Exact Results for Spin Dynamics and Fractionalization in the Kitaev Model, Phys. Rev. Lett. 98, 247201 (2007).
[75]
G. Baskaran, D. Sen, and R. Shankar, Spin- S Kitaev Model: Classical Ground States, Order from Disorder, and Exact Correlation Functions, Phys. Rev. B 78, 115116 (2008).
[76]
Z. Nussinov and G. Ortiz, Bond Algebras and Exact Solvability of Hamiltonians: Spin S=½ Multilayer Systems, Physical Review B 79, 214440 (2009).
[77]
K. O’Brien, M. Hermanns, and S. Trebst, Classification of Gapless Z₂ Spin Liquids in Three-Dimensional Kitaev Models, Phys. Rev. B 93, 085101 (2016).
[78]
M. Hermanns, K. O’Brien, and S. Trebst, Weyl Spin Liquids, Phys. Rev. Lett. 114, 157202 (2015).
[79]
H. Yao and S. A. Kivelson, An Exact Chiral Spin Liquid with Non-Abelian Anyons, Phys. Rev. Lett. 99, 247203 (2007).
[80]
T. Eschmann, P. A. Mishchenko, T. A. Bojesen, Y. Kato, M. Hermanns, Y. Motome, and S. Trebst, Thermodynamics of a Gauge-Frustrated Kitaev Spin Liquid, Phys. Rev. Research 1, 032011 (2019).
[81]
V. Peri, S. Ok, S. S. Tsirkin, T. Neupert, G. Baskaran, M. Greiter, R. Moessner, and R. Thomale, Non-Abelian Chiral Spin Liquid on a Simple Non-Archimedean Lattice, Phys. Rev. B 101, 041114 (2020).
[82]
D. Weaire and M. F. Thorpe, Electronic Properties of an Amorphous Solid. I. A Simple Tight-Binding Theory, Phys. Rev. B 4, 2508 (1971).
[83]
G. P. Betteridge, A Possible Model of Amorphous Silicon and Germanium, J. Phys. C: Solid State Phys. 6, L427 (1973).
[84]
N. P. Mitchell, L. M. Nash, D. Hexner, A. M. Turner, and W. T. M. Irvine, Amorphous topological insulators constructed from random point sets, Nature Phys 14, 380 (2018).
[85]
A. Agarwala, Topological Insulators in Amorphous Systems, in Excursions in Ill-Condensed Quantum Matter: From Amorphous Topological Insulators to Fractional Spins, edited by A. Agarwala (Springer International Publishing, Cham, 2019), pp. 61–79.
[86]
Q. Marsal, D. Varjas, and A. G. Grushin, Topological Weaire-Thorpe Models of Amorphous Matter, Proc. Natl. Acad. Sci. U.S.A. 117, 30260 (2020).
[87]
M. Costa, G. R. Schleder, M. Buongiorno Nardelli, C. Lewenkopf, and A. Fazzio, Toward Realistic Amorphous Topological Insulators, Nano Lett. 19, 8941 (2019).
[88]
A. Agarwala, V. Juričić, and B. Roy, Higher-Order Topological Insulators in Amorphous Solids, Phys. Rev. Research 2, 012067 (2020).
[89]
H. Spring, A. Akhmerov, and D. Varjas, Amorphous Topological Phases Protected by Continuous Rotation Symmetry, SciPost Physics 11, 022 (2021).
[90]
P. Corbae et al., Evidence for Topological Surface States in Amorphous Bi₂Se₃, arXiv Preprint arXiv:1910.13412 (2019).
[91]
[92]
W. McMillan and J. Mochel, Electron Tunneling Experiments on Amorphous Ge_{1- x} Au_x, Phys. Rev. Lett. 46, 556 (1981).
[93]
L. V. Meisel and P. J. Cote, Eliashberg Function in Amorphous Metals, Phys. Rev. B 23, 5834 (1981).
[94]
G. Bergmann, Amorphous Metals and Their Superconductivity, Physics Reports 27, 159 (1976).
[95]
S. Manna, S. K. Das, and B. Roy, Noncrystalline Topological Superconductors, arXiv:2207.02203.
[96]
S. Kim, A. Agarwala, and D. Chowdhury, Fractionalization and Topology in Amorphous Electronic Solids, arXiv Preprint arXiv:2205.11523 (2022).
[97]
A. Aharony, Critical Behavior of Amorphous Magnets, Phys. Rev. B 12, 1038 (1975).
[98]
G. A. Petrakovskii, Amorphous Magnetic Materials, Soviet Physics Uspekhi 24, 511 (1981).
[99]
T. Kaneyoshi, Introduction to Amorphous Magnets (WORLD SCIENTIFIC, 1992).
[100]
T. Kaneyoshi, Amorphous Magnetism (CRC Press, Boca Raton, 2017).
[101]
[102]
J. A. Plascak, L. E. Zamora, and G. A. Pérez Alcazar, Ising Model for Disordered Ferromagnetic Fe-Al Alloys, Phys. Rev. B 61, 3188 (2000).
[103]
J. M. D. Coey, Amorphous Magnetic Order, Journal of Applied Physics 49, 1646 (1978).
[104]
J. Knolle and R. Moessner, A Field Guide to Spin Liquids, Annual Review of Condensed Matter Physics 10, 451 (2019).
[105]
L. Savary and L. Balents, Quantum Spin Liquids: A Review, Rep. Prog. Phys. 80, 016502 (2016).
[106]
C. Lacroix, P. Mendels, and F. Mila, editors, Introduction to Frustrated Magnetism, Vol. 164 (Springer-Verlag, Berlin Heidelberg, 2011).
[107]
T. Eschmann, P. A. Mishchenko, K. O’Brien, T. A. Bojesen, Y. Kato, M. Hermanns, Y. Motome, and S. Trebst, Thermodynamic Classification of Three-Dimensional Kitaev Spin Liquids, Phys. Rev. B 102, 075125 (2020).

  1. The Pauli exclusion principle is special in that it can be treated much more simply that other interactions, this relates to the fact that it is a hard constraint.↩︎

  2. In fact the FQH state can be described within a Ginzburg-Landau like paradigm but it requires a non-local order parameter  [51,52]. Phases like this are said to possess topological order defined by the fact that they cannot be smoothly deformed into a product state  [53,54].↩︎