Spin ice

Last updated
Figure 1. The arrangement of hydrogen atoms (black circles) about oxygen atoms (open circles) in ice. Two hydrogen atoms (bottom ones) are close to the central oxygen atom while two of them (top ones) are far and closer to the two other (top left and top right) oxygen atoms. Water ice.png
Figure 1. The arrangement of hydrogen atoms (black circles) about oxygen atoms (open circles) in ice. Two hydrogen atoms (bottom ones) are close to the central oxygen atom while two of them (top ones) are far and closer to the two other (top left and top right) oxygen atoms.

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. [1] 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).

Contents

Experiments have found evidence for the existence of deconfined magnetic monopoles in these materials, [2] [3] [4] with properties resembling those of the hypothetical magnetic monopoles postulated to exist in vacuum.

Technical description

In 1935, Linus Pauling noted that the hydrogen atoms in water ice would be expected to remain disordered even at absolute zero. That is, even upon cooling to zero temperature, water ice is expected to have residual entropy, i.e., intrinsic randomness. This is due to the fact that the hexagonal crystalline structure of common water ice contains oxygen atoms with four neighboring hydrogen atoms. In ice, for each oxygen atom, two of the neighboring hydrogen atoms are near (forming the traditional H2O molecule), and two are further away (being the hydrogen atoms of two neighboring water molecules). Pauling noted that the number of configurations conforming to this "two-near, two-far" ice rule grows exponentially with the system size, and, therefore, that the zero-temperature entropy of ice was expected to be extensive. [5] Pauling's findings were confirmed by specific heat measurements, though pure crystals of water ice are particularly hard to create.

Figure 2. Portion of a pyrochlore lattice of corner-linked tetrahedra. The magnetic ions (dark blue spheres) sit on a network of tetrahedra linked at their vertices. The other atoms (e.g. Ti and O) making the pyrochlore crystal structure are not displayed. The magnetic moments (light blue arrows) obey the two-in, two out spin ice rule over the whole lattice. The system is thus in a spin ice state. Fragment of pyrochlore lattice in spin ice state.png
Figure 2. Portion of a pyrochlore lattice of corner-linked tetrahedra. The magnetic ions (dark blue spheres) sit on a network of tetrahedra linked at their vertices. The other atoms (e.g. Ti and O) making the pyrochlore crystal structure are not displayed. The magnetic moments (light blue arrows) obey the two-in, two out spin ice rule over the whole lattice. The system is thus in a spin ice state.

Spin ices are materials that consist of regular corner-linked tetrahedra of magnetic ions, each of which has a non-zero magnetic moment, often abridged to "spin", which must satisfy in their low-energy state a "two-in, two-out" rule on each tetrahedron making the crystalline structure (see figure 2). This is highly analogous to the two-near, two far rule in water ice (see figure 1). Just as Pauling showed that the ice rule leads to an extensive entropy in water ice, so does the two-in, two-out rule in the spin ice systems – these exhibit the same residual entropy properties as water ice. Be that as it may, depending on the specific spin ice material, it is generally much easier to create large single crystals of spin ice materials than water ice crystals. Additionally, the ease to induce interaction of the magnetic moments with an external magnetic field in a spin ice system makes the spin ices more suitable than water ice for exploring how the residual entropy can be affected by external influences.

While Philip Anderson had already noted in 1956 [6] the connection between the problem of the frustrated Ising antiferromagnet on a (pyrochlore) lattice of corner-shared tetrahedra and Pauling's water ice problem, real spin ice materials were only discovered forty years later. [7] The first materials identified as spin ices were the pyrochlores Dy2Ti2O7 (dysprosium titanate), Ho2Ti2O7 (holmium titanate). In addition, compelling evidence has been reported that Dy2Sn2O7 (dysprosium stannate) and Ho2Sn2O7 (holmium stannate) are spin ices. [8] These four compounds belong to the family of rare-earth pyrochlore oxides. CdEr2Se4, a spinel in which the magnetic Er3+ ions sit on corner-linked tetrahedra, also displays spin ice behavior. [9]

Spin ice materials are characterized by a random disorder in the orientation of the moment of the magnetic ions, even when the material is at very low temperatures. Alternating current (AC) magnetic susceptibility measurements find evidence for a dynamic freezing of the magnetic moments as the temperature is lowered somewhat below the temperature at which the specific heat displays a maximum. The broad maximum in the heat capacity does not correspond to a phase transition. Rather, the temperature at which the maximum occurs, about 1 K in Dy2Ti2O7, signals a rapid change in the number of tetrahedra where the two-in, two-out rule is violated. Tetrahedra where the rule is violated are sites where the aforementioned monopoles reside. Mathematically, spin ice configurations can be described by closed Eulerian paths. [10] [11]

Spin ices and magnetic monopoles

Figure 3. The orientation of the magnetic moments (light blue arrows) considering a single tetrahedron within the spin ice state, as in figure 2. Here, the magnetic moments obey the two-in, two-out rule: there is as much "magnetization field" going in the tetrahedron (bottom two arrows) as there is going out (top two arrows). The corresponding magnetization field has zero divergence. There is therefore no sink or source of the magnetization inside the tetrahedron, or no monopole. If a thermal fluctuation caused one of the bottom two magnetic moments to flip from "in" to "out", one would then have a 1-in, 3-out configuration; hence an "outflow' of magnetization, hence a positive divergence, that one could assign to a positively charged monopole of charge +Q. Flipping the two bottom magnetic moments would give a 0-in, 4-out configuration, the maximum possible "outflow" (i.e. divergence) of magnetization and, therefore, an associated monopole of charge +2Q. Tetrahedron in ice rule.png
Figure 3. The orientation of the magnetic moments (light blue arrows) considering a single tetrahedron within the spin ice state, as in figure 2. Here, the magnetic moments obey the two-in, two-out rule: there is as much "magnetization field" going in the tetrahedron (bottom two arrows) as there is going out (top two arrows). The corresponding magnetization field has zero divergence. There is therefore no sink or source of the magnetization inside the tetrahedron, or no monopole. If a thermal fluctuation caused one of the bottom two magnetic moments to flip from "in" to "out", one would then have a 1-in, 3-out configuration; hence an "outflow' of magnetization, hence a positive divergence, that one could assign to a positively charged monopole of charge +Q. Flipping the two bottom magnetic moments would give a 0-in, 4-out configuration, the maximum possible "outflow" (i.e. divergence) of magnetization and, therefore, an associated monopole of charge +2Q.

Spin ices are geometrically frustrated magnetic systems. While frustration is usually associated with triangular or tetrahedral arrangements of magnetic moments coupled via antiferromagnetic exchange interactions, as in Anderson's Ising model, [6] spin ices are frustrated ferromagnets. It is the very strong local magnetic anisotropy from the crystal field forcing the magnetic moments to point either in or out of a tetrahedron that renders ferromagnetic interactions frustrated in spin ices. Most importantly, it is the long-range magnetostatic dipole–dipole interaction, and not the nearest-neighbor exchange, that causes the frustration and the consequential two-in, two-out rule that leads to the spin ice phenomenology. [12] [13]

For a tetrahedron in a two-in, two-out state, the magnetization field is divergent-free; there is as much "magnetization intensity" entering a tetrahedron as there is leaving (see figure 3). In such a divergent-free situation, there exists no source or sink for the field. According to Gauss' theorem (also known as Ostrogradsky's theorem), a nonzero divergence of a field is caused, and can be characterized, by a real number called "charge". In the context of spin ice, such charges characterizing the violation of the two-in, two-out magnetic moment orientation rule are the aforementioned monopoles. [2] [3] [4]

In Autumn 2009, researchers reported experimental observation of low-energy quasiparticles resembling the predicted monopoles in spin ice. [2] A single crystal of the dysprosium titanate spin ice candidate was examined in the temperature range of 0.6–2.0 K. Using neutron scattering, the magnetic moments were shown to align in the spin ice material into interwoven tube-like bundles resembling Dirac strings. At the defect formed by the end of each tube, the magnetic field looks like that of a monopole. Using an applied magnetic field, the researchers were able to control the density and orientation of these strings. A description of the heat capacity of the material in terms of an effective gas of these quasiparticles was also presented. [14] [15]

The effective charge of a magnetic monopole, Q (see figure 3) in both the dysprosium and holmium titanate spin ice compounds is approximately Q = 5 μBÅ−1 (Bohr magnetons per angstrom). [2] The elementary magnetic constituents of spin ice are magnetic dipoles, so the emergence of monopoles is an example of the phenomenon of fractionalization.

The microscopic origin of the atomic magnetic moments in magnetic materials is quantum mechanical; the Planck constant enters explicitly in the equation defining the magnetic moment of an electron, along with its charge and its mass. Yet, the magnetic moments in the dysprosium titanate and the holmium titanate spin ice materials are effectively described by classical statistical mechanics, and not quantum statistical mechanics, over the experimentally relevant and reasonably accessible temperature range (between 0.05 K and 2 K) where the spin ice phenomena manifest themselves. Although the weakness of quantum effects in these two compounds is rather unusual, it is believed to be understood. [16] There is current interest in the search of quantum spin ices, [17] materials in which the laws of quantum mechanics now become needed to describe the behavior of the magnetic moments. Magnetic ions other than dysprosium (Dy) and holmium (Ho) are required to generate a quantum spin ice, with praseodymium (Pr), terbium (Tb) and ytterbium (Yb) being possible candidates. [17] [18] One reason for the interest in quantum spin ice is the belief that these systems may harbor a quantum spin liquid , [19] a state of matter where magnetic moments continue to wiggle (fluctuate) down to absolute zero temperature. The theory [20] describing the low-temperature and low-energy properties of quantum spin ice is akin to that of vacuum quantum electrodynamics, or QED. This constitutes an example of the idea of emergence. [21]

Artificial spin ices

Artificial spin ices are metamaterials consisting of coupled nanomagnets arranged on periodic and aperiodic lattices. [22] These systems have enabled the experimental investigation of a variety of phenomena such as frustration, emergent magnetic monopoles, and phase transitions. In addition, artificial spin ices show potential as reprogrammable magnonic crystals and have been studied for their fast dynamics. A variety of geometries have been explored, including quasicrystalline systems and 3D structures, as well as different magnetic materials to modify anisotropies and blocking temperatures.

For example, polymer magnetic composites comprising 2D lattices of droplets of solid-liquid phase change material, with each droplet containing a single magnetic dipole particle, form an artificial spin ice above the droplet melting point, and, after cooling, a spin glass state with low bulk remanence. Spontaneous emergence of 2D magnetic vortices was observed in such spin ices, which vortex geometries were correlated with the external bulk remanence. [23]

Future work in this field includes further developments in fabrication and characterization methods, exploration of new geometries and material combinations, and potential applications in computation, [24] data storage, and reconfigurable microwave circuits. [25] In 2021 a study demonstrated neuromorphic reservoir computing using artificial spin ice, solving a range of computational tasks using the complex magnetic dynamics of the artificial spin ice. [26] In 2022, another studied achieved an artificial kagome spin ice which could potentially be used in the future for novel high-speed computers with low power consumption. [27]

See also

Related Research Articles

<span class="mw-page-title-main">Dysprosium</span> Chemical element, symbol Dy and atomic number 66

Dysprosium is a chemical element; it has symbol Dy and atomic number 66. It is a rare-earth element in the lanthanide series with a metallic silver luster. Dysprosium is never found in nature as a free element, though, like other lanthanides, it is found in various minerals, such as xenotime. Naturally occurring dysprosium is composed of seven isotopes, the most abundant of which is 164Dy.

<span class="mw-page-title-main">Magnetic monopole</span> Hypothetical particle with one magnetic pole

In particle physics, a magnetic monopole is a hypothetical elementary particle that is an isolated magnet with only one magnetic pole. A magnetic monopole would have a net north or south "magnetic charge". Modern interest in the concept stems from particle theories, notably the grand unified and superstring theories, which predict their existence. The known elementary particles that have electric charge are electric monopoles.

<span class="mw-page-title-main">Negative temperature</span> Physical systems hotter than any other

Certain systems can achieve negative thermodynamic temperature; that is, their temperature can be expressed as a negative quantity on the Kelvin or Rankine scales. This should be distinguished from temperatures expressed as negative numbers on non-thermodynamic Celsius or Fahrenheit scales, which are nevertheless higher than absolute zero. A system with a truly negative temperature on the Kelvin scale is hotter than any system with a positive temperature. If a negative-temperature system and a positive-temperature system come in contact, heat will flow from the negative- to the positive-temperature system. A standard example of such a system is population inversion in laser physics.

<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.

In condensed matter physics, the term geometrical frustration refers to a phenomenon where atoms tend to stick to non-trivial positions or where, on a regular crystal lattice, conflicting inter-atomic forces lead to quite complex structures. As a consequence of the frustration in the geometry or in the forces, a plenitude of distinct ground states may result at zero temperature, and usual thermal ordering may be suppressed at higher temperatures. Much studied examples are amorphous materials, glasses, or dilute magnets.

A single-molecule magnet (SMM) is a metal-organic compound that has superparamagnetic behavior below a certain blocking temperature at the molecular scale. In this temperature range, an SMM exhibits magnetic hysteresis of purely molecular origin. In contrast to conventional bulk magnets and molecule-based magnets, collective long-range magnetic ordering of magnetic moments is not necessary.

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

Helimagnetism is a form of magnetic ordering where spins of neighbouring magnetic moments arrange themselves in a spiral or helical pattern, with a characteristic turn angle of somewhere between 0 and 180 degrees. It results from the competition between ferromagnetic and antiferromagnetic exchange interactions. It is possible to view ferromagnetism and antiferromagnetism as helimagnetic structures with characteristic turn angles of 0 and 180 degrees respectively. Helimagnetic order breaks spatial inversion symmetry, as it can be either left-handed or right-handed in nature.

<span class="mw-page-title-main">Dysprosium titanate</span> Chemical compound

Dysprosium titanate (Dy2Ti2O7) is an inorganic compound, a ceramic of the titanate family, with pyrochlore structure.

In statistical mechanics, the ice-type models or six-vertex models are a family of vertex models for crystal lattices with hydrogen bonds. The first such model was introduced by Linus Pauling in 1935 to account for the residual entropy of water ice. Variants have been proposed as models of certain ferroelectric and antiferroelectric crystals.

<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.

In magnetism, a nanomagnet is a nanoscopic scale system that presents spontaneous magnetic order (magnetization) at zero applied magnetic field (remanence).

<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, 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.

<span class="mw-page-title-main">Steven T. Bramwell</span> British scientist

Steven T. Bramwell is a British physicist and chemist who works at the London Centre for Nanotechnology and the Department of Physics and Astronomy, University College London. He is known for his experimental discovery of spin ice with M. J. Harris and his calculation of a critical exponent observed in two-dimensional magnets with P. C. W. Holdsworth. A probability distribution for global quantities in complex systems, the "Bramwell-Holdsworth-Pinton (BHP) distribution", is named after him.

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.

<span class="mw-page-title-main">Holmium titanate</span> Chemical compound

Holmium titanate is an inorganic compound with the chemical formula Ho2Ti2O7.

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.

Shivaji Lal Sondhi is an Indian-born theoretical physicist who is currently the Wykeham Professor of Physics in the Rudolf Peierls Centre for Theoretical Physics at the University of Oxford, known for contributions to the field of quantum condensed matter. He is son of former Lok Sabha MP Manohar Lal Sondhi.

<span class="mw-page-title-main">Dysprosium stannate</span> Chemical compound

Dysprosium stannate (Dy2Sn2O7) is an inorganic compound, a ceramic of the stannate family, with pyrochlore structure.

References

  1. Bramwell, S. T.; Gingras, M. J. P. (2001). "Spin Ice State in Frustrated Magnetic Pyrochlore Materials". Science. 294 (5546): 1495–1501. arXiv: cond-mat/0201427 . Bibcode:2001Sci...294.1495B. doi:10.1126/science.1064761. PMID   11711667. S2CID   9402061.
  2. 1 2 3 4 Castelnovo, C.; Moessner, R.; Sondhi, S. L. (2008-01-03). "Magnetic monopoles in spin ice". Nature. 451 (7174): 42–45. arXiv: 0710.5515 . Bibcode:2008Natur.451...42C. doi:10.1038/nature06433. ISSN   0028-0836. PMID   18172493. S2CID   2399316.
  3. 1 2 Tchernyshyov, Oleg (2008-01-03). "Magnetism: Freedom for the poles". Nature. 451 (7174): 22–23. Bibcode:2008Natur.451...22T. doi:10.1038/451022b. ISSN   0028-0836. PMID   18172484. S2CID   30259694.
  4. 1 2 Gingras, M.J.P. (2009). "Observing Monopoles in a Magnetic Analog of Ice". Science. 326 (5951): 375–376. arXiv: 1005.3557 . doi:10.1126/science.1181510. PMID   19833948. S2CID   31038263.
  5. Pauling, Linus (1935). "The Structure and Entropy of Ice and of Other Crystals with Some Randomness of Atomic Arrangement". Journal of the American Chemical Society. American Chemical Society (ACS). 57 (12): 2680–2684. doi:10.1021/ja01315a102. ISSN   0002-7863.
  6. 1 2 Anderson, P. W. (15 May 1956). "Ordering and Antiferromagnetism in Ferrites". Physical Review. American Physical Society (APS). 102 (4): 1008–1013. Bibcode:1956PhRv..102.1008A. doi:10.1103/physrev.102.1008. ISSN   0031-899X.
  7. Harris, M. J.; Bramwell, S. T.; McMorrow, D. F.; Zeiske, T.; Godfrey, K. W. (29 September 1997). "Geometrical Frustration in the Ferromagnetic Pyrochlore Ho2Ti2O7" (PDF). Physical Review Letters. American Physical Society (APS). 79 (13): 2554–2557. Bibcode:1997PhRvL..79.2554H. doi:10.1103/physrevlett.79.2554. hdl: 20.500.11820/f7958ee9-5fb1-4965-8ab3-70c50216943c . ISSN   0031-9007. S2CID   121002411.
  8. Matsuhira, Kazuyuki; Hinatsu, Yukio; Tenya, Kenichi; Amitsuka, Hiroshi; Sakakibara, Toshiro (15 June 2002). "Low-Temperature Magnetic Properties of Pyrochlore Stannates". Journal of the Physical Society of Japan. Physical Society of Japan. 71 (6): 1576–1582. Bibcode:2002JPSJ...71.1576M. doi:10.1143/jpsj.71.1576. ISSN   0031-9015.
  9. Lago, J.; Živković, I.; Malkin, B. Z.; Rodriguez Fernandez, J.; Ghigna, P.; Dalmas de Réotier, P.; Yaouanc, A.; Rojo, T. (2010-06-15). "CdEr2Se4: A New Erbium Spin Ice System in a Spinel Structure". Physical Review Letters. 104 (24): 247203. Bibcode:2010PhRvL.104x7203L. doi:10.1103/PhysRevLett.104.247203. hdl: 10902/28560 . PMID   20867332.
  10. Schrijver, A. (1983). "Bounds on the number of Eulerian Orientations". Combinatorica. 3 (3): 375–380. doi:10.1007/BF02579193. S2CID   13708977.
  11. Caravelli, F.; Saccone, M.; Nisoli, C. (2021). "On the Degeneracy of Spin Ice Graphs, and its Estimate via the Bethe Permanent". Proc. R. Soc. A. 477 (20210108). arXiv: 2101.12280 . Bibcode:2021RSPSA.47710108C. doi: 10.1098/rspa.2021.0108 . S2CID   231728393.
  12. den Hertog, Byron C.; Gingras, Michel J. P. (10 April 2000). "Dipolar Interactions and Origin of Spin Ice in Ising Pyrochlore Magnets". Physical Review Letters. American Physical Society (APS). 84 (15): 3430–3433. arXiv: cond-mat/0001369 . Bibcode:2000PhRvL..84.3430D. doi:10.1103/physrevlett.84.3430. ISSN   0031-9007. PMID   11019107. S2CID   45435198.
  13. Isakov, S. V.; Moessner, R.; Sondhi, S. L. (14 November 2005). "Why Spin Ice Obeys the Ice Rules". Physical Review Letters. 95 (21): 217201. arXiv: cond-mat/0502137 . Bibcode:2005PhRvL..95u7201I. doi:10.1103/physrevlett.95.217201. ISSN   0031-9007. PMID   16384174. S2CID   30364648.
  14. "Magnetic Monopoles Detected In A Real Magnet For The First Time". Science Daily. 2009-09-04. Retrieved 2009-09-04.
  15. D.J.P. Morris; D.A. Tennant; S.A. Grigera; B. Klemke; C. Castelnovo; R. Moessner; C. Czternasty; M. Meissner; K.C. Rule; J.-U. Hoffmann; K. Kiefer; S. Gerischer; D. Slobinsky & R.S. Perry (2009-09-03). "Dirac Strings and Magnetic Monopoles in Spin Ice Dy2Ti2O7". Science . 326 (5951): 411–4. arXiv: 1011.1174 . Bibcode:2009Sci...326..411M. doi:10.1126/science.1178868. PMID   19729617. S2CID   206522398.
  16. Rau, Jeffrey G.; Gingras, Michel J. P. (2015). "Magnitude of quantum effects in classical spin ices". Physical Review B. 92 (14): 144417. arXiv: 1503.04808 . Bibcode:2015PhRvB..92n4417R. doi:10.1103/PhysRevB.92.144417. S2CID   119153613.
  17. 1 2 Gingras, M. J. P.; McClarty, P. A. (2014-01-01). "Quantum spin ice: a search for gapless quantum spin liquids in pyrochlore magnets". Reports on Progress in Physics. 77 (5): 056501. arXiv: 1311.1817 . Bibcode:2014RPPh...77e6501G. doi:10.1088/0034-4885/77/5/056501. ISSN   0034-4885. PMID   24787264. S2CID   23594100.
  18. Rau, Jeffrey G.; Gingras, Michel J.P. (2019-03-10). "Frustrated Quantum Rare-Earth Pyrochlores". Annual Review of Condensed Matter Physics. 10 (1): 357–386. arXiv: 1806.09638 . Bibcode:2019ARCMP..10..357R. doi:10.1146/annurev-conmatphys-022317-110520. ISSN   1947-5454. S2CID   85498113.
  19. Balents, Leon (2010-03-10). "Spin liquids in frustrated magnets". Nature. 464 (7286): 199–208. Bibcode:2010Natur.464..199B. doi:10.1038/nature08917. ISSN   0028-0836. PMID   20220838. S2CID   4408289.
  20. Hermele, Michael; Fisher, Matthew P. A.; Balents, Leon (2004-02-12). "Pyrochlore photons: The U(1) spin liquid in a S=1/2 three-dimensional frustrated magnet". Physical Review B. 69 (6): 064404. arXiv: cond-mat/0305401 . Bibcode:2004PhRvB..69f4404H. doi:10.1103/PhysRevB.69.064404. S2CID   28840838.
  21. Rehn, J.; Moessner, R. (2016-05-19). "Maxwell electromagnetism as an emergent phenomenon in condensed matter". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 374 (2075): 20160093. arXiv: 1605.05874 . Bibcode:2016RSPTA.37460093R. doi:10.1098/rsta.2016.0093. PMID   27458263. S2CID   206159482.
  22. Skjærvø, Sandra H.; Marrows, Christopher H.; Stamps, Robert L.; Heyderman, Laura J. (8 November 2019). "Advances in artificial spin ice". Nature Reviews Physics. 2 (1): 13–28. doi:10.1038/s42254-019-0118-3. eISSN   2522-5820. S2CID   207959741.
  23. Kaya, Kerem; Iseri, Emre; van der Wijngaart, Wouter (6 December 2022). "Soft metamaterial with programmable ferromagnetism". Microsystems & Nanoengineering. 8 (1): 127. Bibcode:2022MicNa...8..127K. doi:10.1038/s41378-022-00463-2. eISSN   2055-7434. PMC   9722694 . PMID   36483621.
  24. Caravelli, F.; Nisoli, C. (2020). "Logical gates embedding in artificial spin ice". New J. Phys. 22 (103052): 103052. arXiv: 1810.09190 . Bibcode:2020NJPh...22j3052C. doi: 10.1088/1367-2630/abbf21 . S2CID   216056260.
  25. Heyderman, L.J. (2022). "Spin ice devices from nanomagnets". Nature Nanotechnology. 17 (5): 435–436. Bibcode:2022NatNa..17..435H. doi:10.1038/s41565-022-01088-2. PMID   35513585. S2CID   248530509.
  26. Gartside, J. C.; Stenning, K. D.; Vanston, A.; Holder, H.H.; Arroo, D.M.; Dion, T.; Caravelli, F.; Kurebayashi, H.; Branford, W. R. (2022-04-04). "Reconfigurable training and reservoir computing in an artificial spin-vortex ice via spin-wave fingerprinting". Nature Nanotechnology. 17 (5): 460–469. arXiv: 2107.08941 . Bibcode:2022NatNa..17..460G. doi:10.1038/s41565-022-01091-7. PMID   35513584. S2CID   246431279.
  27. "A look into the magnetic future | Our Research | Paul Scherrer Institut (PSI)". www.psi.ch. 2022-04-04. Retrieved 2022-04-10.