Quantum spin liquid

Last updated

In condensed matter physics, a quantum spin liquid is a phase of matter that can be formed by interacting quantum spins in certain magnetic materials. Quantum spin liquids (QSL) are generally characterized by their long-range quantum entanglement, fractionalized excitations, and absence of ordinary magnetic order. [1]

Contents

The quantum spin liquid state was first proposed by physicist Phil Anderson in 1973 as the ground state for a system of spins on a triangular lattice that interact antiferromagnetically with their nearest neighbors, i.e. neighboring spins seek to be aligned in opposite directions. [2] Quantum spin liquids generated further interest when in 1987 Anderson proposed a theory that described high-temperature superconductivity in terms of a disordered spin-liquid state. [3] [4]

Basic properties

Example of a spin liquid emerging from frustrated magnetism

The simplest kind of magnetic phase is a paramagnet, where each individual spin behaves independently of the rest, just like atoms in an ideal gas. This highly disordered phase is the generic state of magnets at high temperatures, where thermal fluctuations dominate. Upon cooling, the spins will often enter a ferromagnet (or antiferromagnet) phase. In this phase, interactions between the spins cause them to align into large-scale patterns, such as domains, stripes, or checkerboards. These long-range patterns are referred to as "magnetic order," and are analogous to the regular crystal structure formed by many solids. [5]

Quantum spin liquids offer a dramatic alternative to this typical behavior. One intuitive description of this state is as a "liquid" of disordered spins, in comparison to a ferromagnetic spin state, [6] much in the way liquid water is in a disordered state compared to crystalline ice. However, unlike other disordered states, a quantum spin liquid state preserves its disorder to very low temperatures. [7] A more modern characterization of quantum spin liquids involves their topological order, [8] long-range quantum entanglement properties, [1] and anyon excitations. [9]

Examples

Several physical models have a disordered ground state that can be described as a quantum spin liquid.

Frustrated magnetic moments

Frustrated Ising spins on a triangle Triangular ising spin.png
Frustrated Ising spins on a triangle

Localized spins are frustrated if there exist competing exchange interactions that can not all be satisfied at the same time, leading to a large degeneracy of the system's ground state. A triangle of Ising spins (meaning that the only possible orientation of the spins are either "up" or "down"), which interact antiferromagnetically, is a simple example for frustration. In the ground state, two of the spins can be antiparallel but the third one cannot. This leads to an increase of possible orientations (six in this case) of the spins in the ground state, enhancing fluctuations and thus suppressing magnetic ordering.

A recent research work used this concept in analyzing brain networks and surprisingly indicated frustrated interactions in the brain corresponding to flexible neural interactions. This observation highlights the generalization of the frustration phenomenon and proposes its investigation in biological systems. [10]

Resonating valence bonds (RVB)

Valence bond solid. The bonds form a specific pattern and consist of pairs of entangled spins. Valence bond solid.png
Valence bond solid. The bonds form a specific pattern and consist of pairs of entangled spins.

To build a ground state without magnetic moment, valence bond states can be used, where two electron spins form a spin 0 singlet due to the antiferromagnetic interaction. If every spin in the system is bound like this, the state of the system as a whole has spin 0 too and is non-magnetic. The two spins forming the bond are maximally entangled, while not being entangled with the other spins. If all spins are distributed to certain localized static bonds, this is called a valence bond solid (VBS).

There are two things that still distinguish a VBS from a spin liquid: First, by ordering the bonds in a certain way, the lattice symmetry is usually broken, which is not the case for a spin liquid. Second, this ground state lacks long-range entanglement. To achieve this, quantum mechanical fluctuations of the valence bonds must be allowed, leading to a ground state consisting of a superposition of many different partitionings of spins into valence bonds. If the partitionings are equally distributed (with the same quantum amplitude), there is no preference for any specific partitioning ("valence bond liquid"). This kind of ground state wavefunction was proposed by P. W. Anderson in 1973 as the ground state of spin liquids [2] and is called a resonating valence bond (RVB) state. These states are of great theoretical interest as they are proposed to play a key role in high-temperature superconductor physics. [4]

Excitations

Spinon moving in spin liquids Spinon moving.png
Spinon moving in spin liquids

The valence bonds do not have to be formed by nearest neighbors only and their distributions may vary in different materials. Ground states with large contributions of long range valence bonds have more low-energy spin excitations, as those valence bonds are easier to break up. On breaking, they form two free spins. Other excitations rearrange the valence bonds, leading to low-energy excitations even for short-range bonds. Something very special about spin liquids is that they support exotic excitations, meaning excitations with fractional quantum numbers. A prominent example is the excitation of spinons which are neutral in charge and carry spin . In spin liquids, a spinon is created if one spin is not paired in a valence bond. It can move by rearranging nearby valence bonds at low energy cost.

Realizations of (stable) RVB states

The first discussion of the RVB state on square lattice using the RVB picture [11] only consider nearest neighbour bonds that connect different sub-lattices. The constructed RVB state is an equal amplitude superposition of all the nearest-neighbour bond configurations. Such a RVB state is believed to contain emergent gapless gauge field which may confine the spinons etc. So the equal-amplitude nearest-neighbour RVB state on square lattice is unstable and does not corresponds to a quantum spin phase. It may describe a critical phase transition point between two stable phases. A version of RVB state which is stable and contains deconfined spinons is the chiral spin state. [12] [13] Later, another version of stable RVB state with deconfined spinons, the Z2 spin liquid, is proposed, [14] [15] which realizes the simplest topological orderZ2 topological order. Both chiral spin state and Z2 spin liquid state have long RVB bonds that connect the same sub-lattice. In chiral spin state, different bond configurations can have complex amplitudes, while in Z2 spin liquid state, different bond configurations only have real amplitudes. The RVB state on triangle lattice also realizes the Z2 spin liquid, [16] where different bond configurations only have real amplitudes. The toric code model is yet another realization of Z2 spin liquid (and Z2 topological order) that explicitly breaks the spin rotation symmetry and is exactly solvable. [17]

Experimental signatures and probes

Since there is no single experimental feature which identifies a material as a spin liquid, several experiments have to be conducted to gain information on different properties which characterize a spin liquid. [18]

Magnetic susceptibility

In a high-temperature, classical paramagnet phase, the magnetic susceptibility is given by the Curie–Weiss law

Fitting experimental data to this equation determines a phenomenological Curie–Weiss temperature, . There is a second temperature, , where magnetic order in the material begins to develop, as evidenced by a non-analytic feature in . The ratio of these is called the frustration parameter

In a classic antiferromagnet, the two temperatures should coincide and give . An ideal quantum spin liquid would not develop magnetic order at any temperature and so would have a diverging frustration parameter . [19] A large value is therefore a good indication of a possible spin liquid phase. Some frustrated materials with different lattice structures and their Curie–Weiss temperature are listed in the table below. [7] All of them are proposed spin liquid candidates.

MaterialLattice
κ-(BEDT-TTF)2Cu2(CN)3anisotropic triangular-375
ZnCu3(OH)6Cl2 (herbertsmithite) Kagome -241
BaCu3V2O8(OH)2 (vesignieite) Kagome
Na4Ir3O8Hyperkagome-650
PbCuTe2O6Hyperkagome-22 [20]
Cu-(1,3-benzenedicarboxylate) Kagome -33 [21]
Rb2Cu3SnF12 Kagome [22]
1T-TaS2Triangular

Other

One of the most direct evidence for absence of magnetic ordering give NMR or μSR experiments. If there is a local magnetic field present, the nuclear or muon spin would be affected which can be measured. 1H-NMR measurements [23] on κ-(BEDT-TTF)2Cu2(CN)3 have shown no sign of magnetic ordering down to 32 mK, which is four orders of magnitude smaller than the coupling constant J≈250 K [24] between neighboring spins in this compound. Further investigations include:

Herbertsmithite, the mineral whose ground state was shown to have QSL behaviour Herbertsmithite-herb03a.jpg
Herbertsmithite, the mineral whose ground state was shown to have QSL behaviour

Candidate materials

RVB type

Neutron scattering measurements of cesium chlorocuprate Cs2CuCl4, a spin-1/2 antiferromagnet on a triangular lattice, displayed diffuse scattering. This was attributed to spinons arising from a 2D RVB state. [26] Later theoretical work challenged this picture, arguing that all experimental results were instead consequences of 1D spinons confined to individual chains. [27]

Afterwards, it was observed in an organic Mott insulator (κ-(BEDT-TTF)2Cu2(CN)3) by Kanoda's group in 2003. [23] It may correspond to a gapless spin liquid with spinon Fermi surface (the so-called uniform RVB state). [2] The peculiar phase diagram of this organic quantum spin liquid compound was first thoroughly mapped using muon spin spectroscopy. [28]

Herbertsmithite

Herbertsmithite is one of the most extensively studied QSL candidate materials. [19] It is a mineral with chemical composition ZnCu3(OH)6Cl2 and a rhombohedral crystal structure. Notably, the copper ions within this structure form stacked two-dimensional layers of kagome lattices. Additionally, superexchange over the oxygen bonds creates a strong antiferromagnetic interaction between the copper spins within a single layer, whereas coupling between layers is negligible. [19] Therefore, it is a good realization of the antiferromagnetic spin-1/2 Heisenberg model on the kagome lattice, which is a prototypical theoretical example of a quantum spin liquid. [29] [30]

Synthetic, polycrystalline herbertsmithite powder was first reported in 2005, and initial magnetic susceptibility studies showed no signs of magnetic order down to 2K. [31] In a subsequent study, the absence of magnetic order was verified down to 50 mK, inelastic neutron scattering measurements revealed a broad spectrum of low energy spin excitations, and low-temperature specific heat measurements had power law scaling. This gave compelling evidence for a spin liquid state with gapless spinon excitations. [32] A broad array of additional experiments, including 17O NMR, [33] and neutron spectroscopy of the dynamic magnetic structure factor, [34] reinforced the identification of herbertsmithite as a gapless spin liquid material, although the exact characterization remained unclear as of 2010. [35]

Large (millimeter size) single crystals of herbertsmithite were grown and characterized in 2011. [36] These enabled more precise measurements of possible spin liquid properties. In particular, momentum-resolved inelastic neutron scattering experiments showed a broad continuum of excitations. This was interpreted as evidence for gapless, fractionalized spinons. [37] Follow-up experiments (using 17O NMR and high-resolution, low-energy neutron scattering) refined this picture and determined there was actually a small spinon excitation gap of 0.07–0.09 meV. [38] [39]

Some measurements were suggestive of quantum critical behavior. [40] [41] Magnetic response of this material displays scaling relation in both the bulk ac susceptibility and the low energy dynamic susceptibility, with the low temperature heat capacity strongly depending on magnetic field. [42] [43] This scaling is seen in certain quantum antiferromagnets, heavy-fermion metals, and two-dimensional 3He as a signature of proximity to a quantum critical point. [44]

In 2020, monodisperse single-crystal nanoparticles of herbertsmithite (~10 nm) were synthesized at room temperature, using gas-diffusion electrocrystallization, showing that their spin liquid nature persists at such small dimensions. [45]

Fig. 1: T-dependence of the electronic specific heat C/T of YbRh2Si2 at different magnetic fields as shown in the legend. The values of (C/T)max and Tmax at B=8 Tesla are shown. The maximum (C/T)max decreases with growing magnetic field B, while Tmax shifts to higher T reaching 14 K at B=18 Tesla. Observing that C/T~kh~M*, one concludes that SCQSL in ZnCu3(OH)6Cl2 shown in Fig. 2 exhibits the similar behavior as heavy fermions in YbRh2Si2. Herb4.png
Fig. 1: T-dependence of the electronic specific heat C/T of YbRh2Si2 at different magnetic fields as shown in the legend. The values of (C/T)max and Tmax at B=8 Tesla are shown. The maximum (C/T)max decreases with growing magnetic field B, while Tmax shifts to higher T reaching 14 K at B=18 Tesla. Observing that C/T~χ~M*, one concludes that SCQSL in ZnCu3(OH)6Cl2 shown in Fig. 2 exhibits the similar behavior as heavy fermions in YbRh2Si2.
Fig.2: T-dependence of the magnetic susceptibility kh at different magnetic fields for ZnCu3(OH)6Cl2. The values of khmax and Tmax at B=7 Tesla are shown. T-dependence T at B=0 is depicted by the solid curve. The maximum khmax(T) decreases as magnetic field B grows, while Tmax(B) shifts to higher T reaching 15 K at B=14 Tesla. Observing that kh~C/T~M*, one concludes that the specific heat of YbRh2Si2 shown in Fig. 1 exhibits the similar behavior as kh does. Thus, SCQSL in ZnCu3(OH)6Cl2 behaves as heavy fermions in YbRh2Si2. Magnetic susceptibility of a function of temperature vs magnetic field - Herbertsmithite.png
Fig.2: T-dependence of the magnetic susceptibility χ at different magnetic fields for ZnCu3(OH)6Cl2. The values of χmax and Tmax at B=7 Tesla are shown. T-dependence T at B=0 is depicted by the solid curve. The maximum χmax(T) decreases as magnetic field B grows, while Tmax(B) shifts to higher T reaching 15 K at B=14 Tesla. Observing that χ~C/T~M*, one concludes that the specific heat of YbRh2Si2 shown in Fig. 1 exhibits the similar behavior as χ does. Thus, SCQSL in ZnCu3(OH)6Cl2 behaves as heavy fermions in YbRh2Si2.

It may realize a U(1)-Dirac spin liquid. [48]

Kitaev spin liquids

Another evidence of quantum spin liquid was observed in a 2-dimensional material in August 2015. The researchers of Oak Ridge National Laboratory, collaborating with physicists from the University of Cambridge, and the Max Planck Institute for the Physics of Complex Systems in Dresden, Germany, measured the first signatures of these fractional particles, known as Majorana fermions, in a two-dimensional material with a structure similar to graphene. Their experimental results successfully matched with one of the main theoretical models for a quantum spin liquid, known as a Kitaev honeycomb model. [49] [50]

Strongly correlated quantum spin liquid

The strongly correlated quantum spin liquid (SCQSL) is a specific realization of a possible quantum spin liquid (QSL) [7] [40] representing a new type of strongly correlated electrical insulator (SCI) that possesses properties of heavy fermion metals with one exception: it resists the flow of electric charge. [47] [51] At low temperatures T the specific heat of this type of insulator is proportional to Tn, with n less or equal 1 rather than n=3, as it should be in the case of a conventional insulator whose heat capacity is proportional to T3. When a magnetic field B is applied to SCI the specific heat depends strongly on B, contrary to conventional insulators. There are a few candidates of SCI; the most promising among them is Herbertsmithite, [51] a mineral with chemical structure ZnCu3(OH)6Cl2.

Kagome type

Ca10Cr7O28 is a frustrated kagome bilayer magnet, which does not develop long-range order even below 1 K, and has a diffuse spectrum of gapless excitations.

Toric code type

In December 2021, the first direct measurement of a quantum spin liquid of the toric code type was reported, [52] [53] it was achieved by two teams: one exploring ground state and anyonic excitations on a quantum processor [54] and the other implementing a theoretical blueprint [55] of atoms on a ruby lattice held with optical tweezers on a quantum simulator. [56]

Specific properties: topological fermion condensation quantum phase transition

The experimental facts collected on heavy fermion (HF) metals and two dimensional Helium-3 demonstrate that the quasiparticle effective mass M* is very large, or even diverges. Topological fermion condensation quantum phase transition (FCQPT) preserves quasiparticles, and forms flat energy band at the Fermi level. The emergence of FCQPT is directly related to the unlimited growth of the effective mass M*. [44] Near FCQPT, M* starts to depend on temperature T, number density x, magnetic field B and other external parameters such as pressure P, etc. In contrast to the Landau paradigm based on the assumption that the effective mass is approximately constant, in the FCQPT theory the effective mass of new quasiparticles strongly depends on T, x, B etc. Therefore, to agree/explain with the numerous experimental facts, extended quasiparticles paradigm based on FCQPT has to be introduced. The main point here is that the well-defined quasiparticles determine the thermodynamic, relaxation, scaling and transport properties of strongly correlated Fermi systems and M* becomes a function of T, x, B, P, etc. The data collected for very different strongly correlated Fermi systems demonstrate universal scaling behavior; in other words distinct materials with strongly correlated fermions unexpectedly turn out to be uniform, thus forming a new state of matter that consists of HF metals, quasicrystals, quantum spin liquid, two-dimensional Helium-3, and compounds exhibiting high-temperature superconductivity. [40] [44]

Applications

Materials supporting quantum spin liquid states may have applications in data storage and memory. [57] In particular, it is possible to realize topological quantum computation by means of spin-liquid states. [58] Developments in quantum spin liquids may also help in the understanding of high temperature superconductivity. [59]

Related Research Articles

<span class="mw-page-title-main">Kondo effect</span> Physical phenomenon due to impurities

In physics, the Kondo effect describes the scattering of conduction electrons in a metal due to magnetic impurities, resulting in a characteristic change i.e. a minimum in electrical resistivity with temperature. The cause of the effect was first explained by Jun Kondo, who applied third-order perturbation theory to the problem to account for scattering of s-orbital conduction electrons off d-orbital electrons localized at impurities. Kondo's calculation predicted that the scattering rate and the resulting part of the resistivity should increase logarithmically as the temperature approaches 0 K. Experiments in the 1960s by Myriam Sarachik at Bell Laboratories provided the first data that confirmed the Kondo effect. Extended to a lattice of magnetic impurities, the Kondo effect likely explains the formation of heavy fermions and Kondo insulators in intermetallic compounds, especially those involving rare earth elements such as cerium, praseodymium, and ytterbium, and actinide elements such as uranium. The Kondo effect has also been observed in quantum dot systems.

In condensed matter physics, a quasiparticle is a concept used to describe a collective behavior of a group of particles that can be treated as if they were a single particle. Formally, quasiparticles and collective excitations are closely related phenomena that arise when a microscopically complicated system such as a solid behaves as if it contained different weakly interacting particles in vacuum.

<span class="mw-page-title-main">Magnon</span> Spin 1 quasiparticle; quantum of a spin wave

A magnon is a quasiparticle, a collective excitation of the spin structure of an electron in a crystal lattice. In the equivalent wave picture of quantum mechanics, a magnon can be viewed as a quantized spin wave. Magnons carry a fixed amount of energy and lattice momentum, and are spin-1, indicating they obey boson behavior.

The fractional quantum Hall effect (FQHE) is a physical phenomenon in which the Hall conductance of 2-dimensional (2D) electrons shows precisely quantized plateaus at fractional values of , where e is the electron charge and h is the Planck constant. It is a property of a collective state in which electrons bind magnetic flux lines to make new quasiparticles, and excitations have a fractional elementary charge and possibly also fractional statistics. The 1998 Nobel Prize in Physics was awarded to Robert Laughlin, Horst Störmer, and Daniel Tsui "for their discovery of a new form of quantum fluid with fractionally charged excitations" The microscopic origin of the FQHE is a major research topic in condensed matter physics.

<span class="mw-page-title-main">Topological order</span> Type of order at absolute zero

In physics, topological order is a kind of order in the zero-temperature phase of matter. Macroscopically, topological order is defined and described by robust ground state degeneracy and quantized non-Abelian geometric phases of degenerate ground states. Microscopically, topological orders correspond to patterns of long-range quantum entanglement. States with different topological orders cannot change into each other without a phase transition.

<span class="mw-page-title-main">Wigner crystal</span> Solid (crystalline) phase of electrons

A Wigner crystal is the solid (crystalline) phase of electrons first predicted by Eugene Wigner in 1934. A gas of electrons moving in a uniform, inert, neutralizing background will crystallize and form a lattice if the electron density is less than a critical value. This is because the potential energy dominates the kinetic energy at low densities, so the detailed spatial arrangement of the electrons becomes important. To minimize the potential energy, the electrons form a bcc lattice in 3D, a triangular lattice in 2D and an evenly spaced lattice in 1D. Most experimentally observed Wigner clusters exist due to the presence of the external confinement, i.e. external potential trap. As a consequence, deviations from the b.c.c or triangular lattice are observed. A crystalline state of the 2D electron gas can also be realized by applying a sufficiently strong magnetic field. However, it is still not clear whether it is the Wigner crystallization that has led to observation of insulating behaviour in magnetotransport measurements on 2D electron systems, since other candidates are present, such as Anderson localization.

<span class="mw-page-title-main">Spin ice</span>

A spin ice is a magnetic substance that does not have a single minimal-energy state. It has magnetic moments (i.e. "spin") as elementary degrees of freedom which are subject to frustrated interactions. By their nature, these interactions prevent the moments from exhibiting a periodic pattern in their orientation down to a temperature much below the energy scale set by the said interactions. Spin ices show low-temperature properties, residual entropy in particular, closely related to those of common crystalline water ice. The most prominent compounds with such properties are dysprosium titanate (Dy2Ti2O7) and holmium titanate (Ho2Ti2O7). The orientation of the magnetic moments in spin ice resembles the positional organization of hydrogen atoms (more accurately, ionized hydrogen, or protons) in conventional water ice (see figure 1).

<span class="mw-page-title-main">Herbertsmithite</span> Halide mineral

Herbertsmithite is a mineral with chemical structure ZnCu3(OH)6Cl2. It is named after the mineralogist Herbert Smith (1872–1953) and was first found in 1972 in Chile. It is polymorphous with kapellasite and closely related to paratacamite. Herbertsmithite is generally found in and around Anarak, Iran, hence its other name, anarakite.

<span class="mw-page-title-main">String-net liquid</span> Condensed matter physics model involving only closed loops

In condensed matter physics, a string-net is an extended object whose collective behavior has been proposed as a physical mechanism for topological order by Michael A. Levin and Xiao-Gang Wen. A particular string-net model may involve only closed loops; or networks of oriented, labeled strings obeying branching rules given by some gauge group; or still more general networks.

Quantum dimer models were introduced to model the physics of resonating valence bond (RVB) states in lattice spin systems. The only degrees of freedom retained from the motivating spin systems are the valence bonds, represented as dimers which live on the lattice bonds. In typical dimer models, the dimers do not overlap.

<span class="mw-page-title-main">Xiao-Gang Wen</span> Chinese-American physicist

Xiao-Gang Wen is a Chinese-American physicist. He is a Cecil and Ida Green Professor of Physics at the Massachusetts Institute of Technology and Distinguished Visiting Research Chair at the Perimeter Institute for Theoretical Physics. His expertise is in condensed matter theory in strongly correlated electronic systems. In Oct. 2016, he was awarded the Oliver E. Buckley Condensed Matter Prize.

<span class="mw-page-title-main">Subir Sachdev</span> Indian physicist

Subir Sachdev is Herchel Smith Professor of Physics at Harvard University specializing in condensed matter. He was elected to the U.S. National Academy of Sciences in 2014, received the Lars Onsager Prize from the American Physical Society and the Dirac Medal from the ICTP in 2018, and was elected Foreign Member of the Royal Society ForMemRS in 2023. He was a co-editor of the Annual Review of Condensed Matter Physics 2017–2019, and is Editor-in-Chief of Reports on Progress in Physics 2022-.

<span class="mw-page-title-main">Piers Coleman</span> British-American physicist

Piers Coleman is a British-born theoretical physicist, working in the field of theoretical condensed matter physics. Coleman is professor of physics at Rutgers University in New Jersey and at Royal Holloway, University of London.

In condensed matter physics, an AKLT model, also known as an Affleck-Kennedy-Lieb-Tasaki model is an extension of the one-dimensional quantum Heisenberg spin model. The proposal and exact solution of this model by Ian Affleck, Elliott H. Lieb, Tom Kennedy and Hal Tasaki provided crucial insight into the physics of the spin-1 Heisenberg chain. It has also served as a useful example for such concepts as valence bond solid order, symmetry-protected topological order and matrix product state wavefunctions.

Roderich Moessner is a theoretical physicist at the Max Planck Institute for the Physics of Complex Systems in Dresden, Germany. His research interests are in condensed matter and materials physics, especially concerning new and topological forms of order, as well as the study of classical and quantum many-body dynamics in and out of equilibrium.

In condensed matter physics, the resonating valence bond theory (RVB) is a theoretical model that attempts to describe high-temperature superconductivity, and in particular the superconductivity in cuprate compounds. It was first proposed by an American physicist P. W. Anderson and Indian theoretical physicist Ganapathy Baskaran in 1987. The theory states that in copper oxide lattices, electrons from neighboring copper atoms interact to form a valence bond, which locks them in place. However, with doping, these electrons can act as mobile Cooper pairs and are able to superconduct. Anderson observed in his 1987 paper that the origins of superconductivity in doped cuprates was in the Mott insulator nature of crystalline copper oxide. RVB builds on the Hubbard and t-J models used in the study of strongly correlated materials.

In solid-state physics, the kagome metal or kagome magnet is a type of ferromagnetic quantum material. The atomic lattice in a kagome magnet has layered overlapping triangles and large hexagonal voids, akin to the kagome pattern in traditional Japanese basket-weaving. This geometry induces a flat electronic band structure with Dirac crossings, in which the low-energy electron dynamics correlate strongly.

In condensed matter physics, the quantum dimer magnet state is one in which quantum spins in a magnetic structure entangle to form a singlet state. These entangled spins act as bosons and their excited states (triplons) can undergo Bose-Einstein condensation (BEC). The quantum dimer system was originally proposed by Matsubara and Matsuda as a mapping of the lattice Bose gas to the quantum antiferromagnet. Quantum dimer magnets are often confused as valence bond solids; however, a valence bond solid requires the breaking of translational symmetry and the dimerizing of spins. In contrast, quantum dimer magnets exist in crystal structures where the translational symmetry is inherently broken. There are two types of quantum dimer models: the XXZ model and the weakly-coupled dimer model. The main difference is the regime in which BEC can occur. For the XXZ model, the BEC occurs upon cooling without a magnetic field and manifests itself as a symmetric dome in the field versus temperature phase diagram centered about H = 0. The weakly-coupled dimer model does not magnetically order in zero magnetic field, but instead orders upon the closing of the spin gap, where the BEC regime begins and is a dome centered at non-zero field.

Elbio Rubén Dagotto is an Argentinian-American theoretical physicist and academic. He is a distinguished professor in the department of physics and astronomy at the University of Tennessee, Knoxville, and Distinguished Scientist in the Materials Science and Technology Division at the Oak Ridge National Laboratory.

The general term mimicry of spin-liquid-by-disorder or appearance of spin-liquid-by-disorder in solid state physics describes the observation of a quantum spin liquid-like (local) state induced by disorder. Laboratory slang calls it just spin-liquid-by-disorder, however, the prefix mimicry or appearance is necessary. The general description is introduced and discussed in the PhD thesis Low-Energy Spin Dynamics in geometrically frustrated 3d-Magnets and Single-Ion Spin Systems by S. A. Bräuninger founded by the SFB 1143. Here, the nature of this observation may have the properity of a local effect and reflects not necessarily the global magnetic ground state. Therefore, the effect is discussed to be a spin-liquid-like state emphasizing a possible local nature. In contrast, order-by-disorder describes a long-range magnetic ordered state which is selected by disorder and associated with a global magnetic ground state of the crystal, e.g. Er2Ti2O7. Therefore, the prefix mimicry or appearance should be used for clarity from a theoretical point of view. Experimentally, the introduction of this term and description is reasonable because of the intensive observations in treatments by selected techniques of local probes like nuclear magnetic resonance (NMR) and muon spin spectroscopy (µSR). However, theoretically, the amount of studies increases, see jammed spin liquid. Such a mimicry is observed and discussed in various works such as disorder-induced spin-liquid-like behavior in kagome-lattice compounds, e.g., Tm3Sb3Zn2O14 and Tm3Sb3Mg2O14, Pr2Zr2O7, Tb2Hf2O7, BaTi0.5Mn0.5O3.

References

  1. 1 2 Savary, L.; Balents, L. (2017). "Quantum spin liquids: a review". Reports on Progress in Physics. 80 (1): 016502. arXiv: 1601.03742 . Bibcode:2017RPPh...80a6502S. doi:10.1088/0034-4885/80/1/016502. PMID   27823986. S2CID   22285828.
  2. 1 2 3 P. W. Anderson (1973). "Resonating valence bonds: A new kind of insulator?". Materials Research Bulletin. 8 (2): 153–160. doi:10.1016/0025-5408(73)90167-0.
  3. Trafton, Anne (March 28, 2011). "A new spin on superconductivity?". MIT News. Retrieved 24 December 2012.
  4. 1 2 P. W. Anderson (1987). "The resonating valence bond state in La2CuO4 and superconductivity". Science. 235 (4793): 1196–1198. Bibcode:1987Sci...235.1196A. doi:10.1126/science.235.4793.1196. PMID   17818979. S2CID   28146486.
  5. Chaikin, Paul M; Lubensky, Tom C (1995). Principles of Condensed-Matter Physics . Cambridge university press. ISBN   9780521432245.
  6. Wilkins, Alasdair (August 15, 2011). "A Strange New Quantum State of Matter: Spin Liquids". io9. Retrieved 23 December 2012.
  7. 1 2 3 Leon Balents (2010). "Spin liquids in frustrated magnets". Nature. 464 (7286): 199–208. Bibcode:2010Natur.464..199B. doi:10.1038/nature08917. PMID   20220838. S2CID   4408289.
  8. Wolchover, Natalie (2018-01-03). "Physicists Aim to Classify All Possible Phases of Matter". Quanta Magazine . Retrieved 2019-05-05.
  9. Wilczek, Frank (2017-02-28). "Inside the Knotty World of 'Anyon' Particles". Quanta Magazine . Retrieved 2019-05-05.
  10. Saberi M, Khosrowabadi R, Khatibi A, Misic B, Jafari G (October 2022). "Pattern of frustration formation in the functional brain network". Network Neuroscience. 6 (4): 1334–1356. doi: 10.1162/netn_a_00268 .
  11. Kivelson, Steven A.; Rokhsar, Daniel S.; Sethna, James P. (1987). "Topology of the resonating valence-bond state: Solitons and high-Tc superconductivity". Physical Review B. 35 (16): 8865–8868. Bibcode:1987PhRvB..35.8865K. doi:10.1103/physrevb.35.8865. PMID   9941277.
  12. Kalmeyer, V.; Laughlin, R. B. (1987). "Equivalence of the resonating-valence-bond and fractional quantum Hall states". Physical Review Letters. 59 (18): 2095–2098. Bibcode:1987PhRvL..59.2095K. doi:10.1103/physrevlett.59.2095. PMID   10035416.
  13. Wen, Xiao-Gang; Wilczek, F.; Zee, A. (1989). "Chiral Spin States and Superconductivity". Physical Review B. 39 (16): 11413–11423. Bibcode:1989PhRvB..3911413W. CiteSeerX   10.1.1.676.519 . doi:10.1103/physrevb.39.11413. PMID   9947970.
  14. Read, N.; Sachdev, Subir (1991). "Large-N expansion for frustrated quantum antiferromagnets". Physical Review Letters. 66 (13): 1773–1776. Bibcode:1991PhRvL..66.1773R. doi:10.1103/physrevlett.66.1773. PMID   10043303.
  15. Wen, Xiao-Gang (1991). "Mean Field Theory of Spin Liquid States with Finite Energy Gaps". Physical Review B. 44 (6): 2664–2672. Bibcode:1991PhRvB..44.2664W. doi:10.1103/physrevb.44.2664. PMID   9999836.
  16. Moessner, R.; Sondhi, S. L. (2002). "Resonating Valence Bond Liquid Physics on the Triangular Lattice". Progress of Theoretical Physics Supplement. 145: 37–42. arXiv: cond-mat/0205029 . Bibcode:2002PThPS.145...37M. doi:10.1143/PTPS.145.37. S2CID   119370249.
  17. Kitaev, A.Yu.; Balents, Leon (2003). "Fault-tolerant quantum computation by anyons". Annals of Physics. 303 (1): 2–30. arXiv: quant-ph/9707021 . Bibcode:2003AnPhy.303....2K. doi:10.1016/S0003-4916(02)00018-0. S2CID   119087885.
  18. Knolle, Johannes; Moessner, Roderich (2019). "A field guide to spin liquids". Annual Review of Condensed Matter Physics. 10: 451–472. arXiv: 1804.02037 . Bibcode:2019ARCMP..10..451K. doi:10.1146/annurev-conmatphys-031218-013401. S2CID   85529148.
  19. 1 2 3 Norman, M.R. (2016). "Herbertsmithite and the Search for the Quantum Spin Liquid". Reviews of Modern Physics. 88 (4): 041002. arXiv: 1710.02991 . doi:10.1103/RevModPhys.88.041002. S2CID   118727125.
  20. Phys. Rev. Lett. 116, 107203 (2016)
  21. Nytko, Emily A.; Helton, Joel S.; Müller, Peter; Nocera, Daniel G. (2008). "A Structurally Perfect S = 1/2 Metal−Organic Hybrid Kagome Antiferromagnet". Journal of the American Chemical Society. 130 (10): 2922–2923. doi:10.1021/ja709991u. PMID   18275194.
  22. Matan, K.; Ono, T.; Fukumoto, Y.; Sato, T. J.; et al. (2010). "Pinwheel valence-bond solid and triplet excitations in the two-dimensional deformed kagome lattice". Nature Physics. 6 (11): 865–869. arXiv: 1007.3625 . Bibcode:2010NatPh...6..865M. doi:10.1038/nphys1761. S2CID   118664640.
  23. 1 2 Y. Shimizu; K. Miyagawa; K. Kanoda; M. Maesato; et al. (2003). "Spin Liquid State in an Organic Mott Insulator with a Triangular Lattice". Physical Review Letters. 91 (10): 107001. arXiv: cond-mat/0307483 . Bibcode:2003PhRvL..91j7001S. doi:10.1103/PhysRevLett.91.107001. PMID   14525498. S2CID   4652670.
  24. In literature, the value of J is commonly given in units of temperature () instead of energy.
  25. T. Ng & P. A. Lee (2007). "Power-Law Conductivity inside the Mott Gap: Application to κ-(BEDT-TTF)2Cu2(CN)3". Physical Review Letters. 99 (15): 156402. arXiv: 0706.0050 . Bibcode:2007PhRvL..99o6402N. doi:10.1103/PhysRevLett.99.156402. PMID   17995193. S2CID   45188091.
  26. Coldea, R.; Tennant, D.A.; Tsvelik, A.M.; Tylczynski, Z. (12 Feb 2001). "Experimental realization of a 2D fractional quantum spin liquid". Physical Review Letters. 86 (7): 1335–1338. arXiv: cond-mat/0007172 . Bibcode:2001PhRvL..86.1335C. doi:10.1103/PhysRevLett.86.1335. PMID   11178077. S2CID   39524266. Note that the preprint was uploaded in 2000.
  27. Kohno, Masanori; Starkh, Oleg A; Balents, Leon (2007). "Spinons and triplons in spatially anisotropic frustrated antiferromagnets". Nature Physics. 3 (11): 790. arXiv: 0706.2012 . Bibcode:2007NatPh...3..790K. doi:10.1038/nphys749. S2CID   28004603.
  28. Pratt, F. L.; Baker, P. J.; Blundell, S. J.; Lancaster, T.; et al. (2011). "Magnetic and non-magnetic phases of a quantum spin liquid". Nature. 471 (7340): 612–616. Bibcode:2011Natur.471..612P. doi:10.1038/nature09910. PMID   21455176. S2CID   4430673.
  29. Elser, Veit (1989). "Nuclear antiferromagnetism in a registered 3He solid". Physical Review Letters. 62 (20): 2405–2408. Bibcode:1989PhRvL..62.2405E. doi:10.1103/PhysRevLett.62.2405. PMID   10039977.
  30. Yan, Simeng; Huse, David A; White, Steven R (2011). "Spin-liquid ground state of the S=1/2 kagome Heisenberg antiferromagnet". Science. 332 (6034): 1173–1176. arXiv: 1011.6114 . Bibcode:2011Sci...332.1173Y. doi:10.1126/science.1201080. PMID   21527676. S2CID   34864628.
  31. Shores, Matthew P; Nytko, Emily A; Bartlett, Bart M; Nocera, Daniel G (2005). "A Structurally Perfect S=1/2 Kagome Antigerromagnet". Journal of the American Chemical Society. 127 (39): 13462–13463. doi:10.1021/ja053891p. PMID   16190686.
  32. Helton, J. S.; et al. (2007). "Spin Dynamics of the Spin-1/2 Kagome Lattice Antiferromagnet ZnCu3(OH)6Cl2". Physical Review Letters . 98 (10): 107204. arXiv: cond-mat/0610539 . Bibcode:2007PhRvL..98j7204H. doi:10.1103/PhysRevLett.98.107204. PMID   17358563. S2CID   23174611.
  33. Olariu, A; et al. (2008). "17O NMR Study of the Intrinsic Magnetic Susceptibility and Spin Dynamics of the Quantum Kagome Antiferromagnet ZnCu3(OH)6Cl2". Physical Review Letters. 100 (9): 087202. arXiv: 0711.2459 . Bibcode:2008PhRvL.100h7202O. doi:10.1103/PhysRevLett.100.087202. PMID   18352658. S2CID   2682652.
  34. de Vries, M. A.; Stewart, J. R.; Deen, P. P.; Piatek, J. O.; Nilsen, G. J.; Ronnow, H. M.; Harrison, A. (2009). "Scale-free antiferromagnetic fluctuations in the S=1/2 kagome antiferromagnet herbertsmithite". Physical Review Letters. 103 (23): 237201. arXiv: 0902.3194 . Bibcode:2009PhRvL.103w7201D. doi:10.1103/PhysRevLett.103.237201. ISSN   0031-9007. PMID   20366167. S2CID   2540295.
  35. Mendels, Philippe; Bert, Fabrice (2010). "Quantum kagome antiferromagnet: ZnCu3(OH)6Cl2". Journal of the Physical Society of Japan. 79 (1): 011001. arXiv: 1001.0801 . Bibcode:2010JPSJ...79a1001M. doi:10.1143/JPSJ.79.011001. S2CID   118545779.
  36. Han, TH; Helton, JS; Chu, S; Prodi, Andrea; Singh, DK; Mazzoli, Claudio; Müller, P; Nocera, DG; Lee, Young S (2011). "Synthesis and characterization of single crystals of the spin-1/2 kagome-lattice antiferromagnets Znx Cu4-x(OH)6Cl2" (PDF). Physical Review B. 83 (10): 100402. arXiv: 1102.2179 . Bibcode:2011PhRvB..83j0402H. doi: 10.1103/PhysRevB.83.100402 .
  37. Han, Tian-Heng; Helton, Joel S; Chu, Shaoyan; Nocera, Daniel G; Rodriguez-Rivera, Jose A; Broholm, Collin; Lee, Young S (2012). "Fractionalized excitations in the spin-liquid state of a kagome-lattice antiferromagnet". Nature. 492 (7429): 406–410. arXiv: 1307.5047 . Bibcode:2012Natur.492..406H. doi:10.1038/nature11659. PMID   23257883. S2CID   4344923.
  38. Fu, Mingxuan; Imai, Takashi; Lee, Young S (2015). "Evidence for a gapped spin-liquid ground state in a kagome Heisenberg antiferromagnet". Science. 350 (6261): 655–658. arXiv: 1511.02174 . Bibcode:2015Sci...350..655F. doi:10.1126/science.aab2120. PMID   26542565. S2CID   22287797.
  39. Han, Tian-Heng; Norman, MR; Wen, J-J; Rodriguez-Rivera, Jose A; Helton, Joel S; Broholm, Collin; Lee, Young S (2016). "Correlated impurities and intrinsic spin-liquid physics in the kagome material herbertsmithite". Physical Review B. 94 (6): 060409. arXiv: 1512.06807 . Bibcode:2016PhRvB..94f0409H. doi:10.1103/PhysRevB.94.060409. S2CID   115149342.
  40. 1 2 3 Amusia, M.; Popov, K.; Shaginyan, V.; Stephanovich, V. (2014). Theory of Heavy-Fermion Compounds - Theory of Strongly Correlated Fermi-Systems. Springer Series in Solid-State Sciences. Vol. 182. Springer. doi:10.1007/978-3-319-10825-4. ISBN   978-3-319-10825-4.
  41. Wen, Jinsheng; Yu, Shun-Li; Li, Shiyan; Yu, Weiqiang; Li, Jian-Xin (12 September 2019). "Experimental identification of quantum spin liquids". npj Quantum Materials. 4 (1): 12. arXiv: 1904.04435 . Bibcode:2019npjQM...4...12W. doi:10.1038/s41535-019-0151-6. ISSN   2397-4648. S2CID   104292206.
  42. 1 2 Helton, J. S.; et al. (2010). "Dynamic Scaling in the Susceptibility of the Spin-1/2 Kagome Lattice Antiferromagnet Herbertsmithite". Physical Review Letters . 104 (14): 147201. arXiv: 1002.1091 . Bibcode:2010PhRvL.104n7201H. doi:10.1103/PhysRevLett.104.147201. PMID   20481955. S2CID   10718733.
  43. de Vries, M. A.; et al. (2008). "The magnetic ground state of an experimental S=1/2 kagomé antiferromagnet". Physical Review Letters . 100 (15): 157205. arXiv: 0705.0654 . Bibcode:2008PhRvL.100o7205D. doi:10.1103/PhysRevLett.100.157205. PMID   18518149. S2CID   118805305.
  44. 1 2 3 Shaginyan, V. R.; Amusia, M. Ya.; Msezane, A. Z.; Popov, K. G. (2010). "Scaling Behavior of Heavy Fermion Metals". Physics Reports . 492 (2–3): 31. arXiv: 1006.2658 . Bibcode:2010PhR...492...31S. CiteSeerX   10.1.1.749.3376 . doi:10.1016/j.physrep.2010.03.001. S2CID   119235769.
  45. Pozo, Guillermo; de la Presa, Patricia; Prato, Rafael; Morales, Irene; Marin, Pilar; Fransaer, Jan; Dominguez-Benetton, Xochitl (2020). "Spin transition nanoparticles made electrochemically". Nanoscale. 12 (9): 5412–5421. doi: 10.1039/C9NR09884D . PMID   32080699.
  46. Gegenwart, P.; et al. (2006). "High-field phase diagram of the heavy-fermion metal YbRh2Si2". New Journal of Physics . 8 (9): 171. Bibcode:2006NJPh....8..171G. doi: 10.1088/1367-2630/8/9/171 .
  47. 1 2 Shaginyan, V. R.; Msezane, A.; Popov, K. (2011). "Thermodynamic Properties of Kagome Lattice in ZnCu3(OH)6Cl2 Herbertsmithite". Physical Review B . 84 (6): 060401. arXiv: 1103.2353 . Bibcode:2011PhRvB..84f0401S. doi:10.1103/PhysRevB.84.060401. S2CID   118651738.
  48. Ying Ran, Michael Hermele, Patrick A. Lee, Xiao-Gang Wen, (2006), "Projected wavefunction study of Spin-1/2 Heisenberg model on the Kagome lattice", https://arxiv.org/abs/cond-mat/0611414
  49. "New state of matter detected in a two-dimensional material" . Retrieved 5 April 2016.
  50. Banerjee, A.; Bridges, C. A.; Yan, J.-Q.; Aczel, A. A.; Li, L.; Stone, M. B.; Granroth, G. E.; Lumsden, M. D.; Yiu, Y.; Knolle, J.; Bhattacharjee, S.; Kovrizhin, D. L.; Moessner, R.; Tennant, D. A.; Mandrus, D. G.; Nagler, S. E. (2016). "Proximate Kitaev quantum spin liquid behaviour in a honeycomb magnet". Nature Materials. 15 (7): 733–740. arXiv: 1504.08037 . Bibcode:2016NatMa..15..733B. doi:10.1038/nmat4604. PMID   27043779. S2CID   3406627.
  51. 1 2 Shaginyan, V. R.; et al. (2012). "Identification of Strongly Correlated Spin Liquid in Herbertsmithite". EPL . 97 (5): 56001. arXiv: 1111.0179 . Bibcode:2012EL.....9756001S. doi:10.1209/0295-5075/97/56001. S2CID   119288349.
  52. Wood, Charlie (2021-12-02). "Realizing topologically ordered states on a quantum processor". Science. 374 (6572): 1237–1241. arXiv: 2104.01180 . Bibcode:2021Sci...374.1237S. doi:10.1126/science.abi8378. PMID   34855491. S2CID   233025160 . Retrieved 2021-12-04.
  53. Wood, Charlie (2021-12-02). "Quantum Simulators Create a Totally New Phase of Matter". Quanta Magazine. Retrieved 2022-03-11.
  54. Satzinger, K. J.; Liu, Y.-J; Smith, A.; Knapp, C.; Newman, M.; Jones, C.; Chen, Z.; Quintana, C.; Mi, X.; Dunsworth, A.; Gidney, C. (2021-12-03). "Realizing topologically ordered states on a quantum processor". Science. 374 (6572): 1237–1241. arXiv: 2104.01180 . Bibcode:2021Sci...374.1237S. doi:10.1126/science.abi8378. PMID   34855491. S2CID   233025160.
  55. Verresen, Ruben; Lukin, Mikhail D.; Vishwanath, Ashvin (2021-07-08). "Prediction of Toric Code Topological Order from Rydberg Blockade". Physical Review X. 11 (3): 031005. arXiv: 2011.12310 . Bibcode:2021PhRvX..11c1005V. doi:10.1103/PhysRevX.11.031005. S2CID   227162637.
  56. Semeghini, G.; Levine, H.; Keesling, A.; Ebadi, S.; Wang, T. T.; Bluvstein, D.; Verresen, R.; Pichler, H.; Kalinowski, M.; Samajdar, R.; Omran, A. (2021-12-03). "Probing topological spin liquids on a programmable quantum simulator". Science. 374 (6572): 1242–1247. arXiv: 2104.04119 . Bibcode:2021Sci...374.1242S. doi:10.1126/science.abi8794. PMID   34855494. S2CID   233204440.
  57. Aguilar, Mario (December 20, 2012). "This Weird Crystal Demonstrates a New Magnetic Behavior That Works Like Magic". Gizmodo. Retrieved 24 December 2012.
  58. Fendley, Paul. "Topological Quantum Computation from non-abelian anyons" (PDF). University of Virginia. Archived from the original (PDF) on 2013-07-28. Retrieved 24 December 2012.
  59. Chandler, David (December 20, 2012). "New kind of magnetism discovered: Experiments demonstrate 'quantum spin liquid'". Phys.org. Retrieved 24 December 2012.