Quantum spin liquid

Source: Wikipedia, the free encyclopedia.

In

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]

The quantum spin liquid state was first proposed by physicist

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

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

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

Localized spins are

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.

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]

  • One possible short-range pairing of spins in a RVB state.
    One possible short-range pairing of spins in a RVB state.
  • Long-range pairing of spins.
    Long-range pairing of spins.

Excitations

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 soluble.[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.

Material Lattice
κ-(BEDT-TTF)2Cu2(CN)3 anisotropic triangular -375
ZnCu3(OH)6Cl2 (herbertsmithite)
Kagome
-241
BaCu3V2O8(OH)2 (vesignieite)
Kagome
Na4Ir3O8 Hyperkagome -650
PbCuTe2O6 Hyperkagome -22[20]
Cu-(1,3-benzenedicarboxylate)
Kagome
-33[21]
Rb2Cu3SnF12
Kagome
[22]
1T-TaS2 Triangular

Other

One of the most direct evidence for absence of magnetic ordering give

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

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

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.

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
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.

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

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[46] 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 χ at different magnetic fields for ZnCu3(OH)6Cl2.[42] The values of χmax and Tmax at B=7 Tesla are shown. T-dependence T(-2/3) 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.[47]

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)

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,

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

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

Applications

Materials supporting quantum spin liquid states may have applications in data storage and memory.

References

  1. ^ a b Savary, L.; Balents, L. (2017). "Quantum spin liquids: a review". Reports on Progress in Physics. 80 (1): 016502.
    S2CID 22285828
    .
  2. ^ .
  3. ^ Trafton, Anne (March 28, 2011). "A new spin on superconductivity?". MIT News. Retrieved 24 December 2012.
  4. ^
    S2CID 28146486
    .
  5. .
  6. ^ Wilkins, Alasdair (August 15, 2011). "A Strange New Quantum State of Matter: Spin Liquids". io9. Retrieved 23 December 2012.
  7. ^
    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. .
  11. .
  12. .
  13. .
  14. .
  15. .
  16. .
  17. .
  18. .
  19. ^ .
  20. ^ Phys. Rev. Lett. 116, 107203 (2016)
  21. PMID 18275194
    .
  22. .
  23. ^ .
  24. ^ In literature, the value of J is commonly given in units of temperature () instead of energy.
  25. S2CID 45188091
    .
  26. . Note that the preprint was uploaded in 2000.
  27. .
  28. .
  29. .
  30. .
  31. .
  32. ^ Helton, J. S.; et al. (2007). "Spin Dynamics of the Spin-1/2 Kagome Lattice Antiferromagnet ZnCu3(OH)6Cl2".
    S2CID 23174611
    .
  33. .
  34. .
  35. .
  36. .
  37. .
  38. .
  39. .
  40. ^ .
  41. .
  42. ^ a b Helton, J. S.; et al. (2010). "Dynamic Scaling in the Susceptibility of the Spin-1/2 Kagome Lattice Antiferromagnet Herbertsmithite".
    S2CID 10718733
    .
  43. ^ de Vries, M. A.; et al. (2008). "The magnetic ground state of an experimental S=1/2 kagomé antiferromagnet".
    S2CID 118805305
    .
  44. ^ a b c Shaginyan, V. R.; Amusia, M. Ya.; Msezane, A. Z.; Popov, K. G. (2010). "Scaling Behavior of Heavy Fermion Metals".
    S2CID 119235769
    .
  45. .
  46. ^ Gegenwart, P.; et al. (2006). "High-field phase diagram of the heavy-fermion metal YbRh2Si2". .
  47. ^ a b Shaginyan, V. R.; Msezane, A.; Popov, K. (2011). "Thermodynamic Properties of Kagome Lattice in ZnCu3(OH)6Cl2 Herbertsmithite".
    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. S2CID 3406627
    .
  51. ^ .
  52. . 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. S2CID 233025160
    .
  55. .
  56. .
  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.