Composite fermion

Last updated

A composite fermion is the topological bound state of an electron and an even number of quantized vortices, sometimes visually pictured as the bound state of an electron and, attached, an even number of magnetic flux quanta. [1] [2] [3] Composite fermions were originally envisioned in the context of the fractional quantum Hall effect, [4] but subsequently took on a life of their own, exhibiting many other consequences and phenomena.

Contents

Vortices are an example of topological defect, and also occur in other situations. Quantized vortices are found in type II superconductors, called Abrikosov vortices. Classical vortices are relevant to the Berezenskii–Kosterlitz–Thouless transition in two-dimensional XY model.

Description

When electrons are confined to two dimensions, cooled to very low temperatures, and subjected to a strong magnetic field, their kinetic energy is quenched due to Landau level quantization. Their behavior under such conditions is governed by the Coulomb repulsion alone, and they produce a strongly correlated quantum liquid. Experiments have shown [1] [2] [3] that electrons minimize their interaction by capturing quantized vortices to become composite fermions. [5] The interaction between composite fermions themselves is often negligible to a good approximation, which makes them the physical quasiparticles of this quantum liquid.

The signature quality of composite fermions, which is responsible for the otherwise unexpected behavior of this system, is that they experience a much smaller magnetic field than electrons. The magnetic field seen by composite fermions is given by

where is the external magnetic field, is the number of vortices bound to composite fermion (also called the vorticity or the vortex charge of the composite fermion), is the particle density in two dimensions, and is called the "flux quantum" (which differs from the superconducting flux quantum by a factor of two). The effective magnetic field is a direct manifestation of the existence of composite fermions, and also embodies a fundamental distinction between electrons and composite fermions.

Sometimes it is said that electrons "swallow" flux quanta each to transform into composite fermions, and the composite fermions then experience the residual magnetic field More accurately, the vortices bound to electrons produce their own geometric phases which partly cancel the Aharonov–Bohm phase due to the external magnetic field to generate a net geometric phase that can be modeled as an Aharonov–Bohm phase in an effective magnetic field

The behavior of composite fermions is similar to that of electrons in an effective magnetic field Electrons form Landau levels in a magnetic field, and the number of filled Landau levels is called the filling factor, given by the expression Composite fermions form Landau-like levels in the effective magnetic field which are called composite fermion Landau levels or levels. One defines the filling factor for composite fermions as This gives the following relation between the electron and composite fermion filling factors

The minus sign occurs when the effective magnetic field is antiparallel to the applied magnetic field, which happens when the geometric phase from the vortices overcompensate the Aharonov–Bohm phase.

Experimental manifestations

The central statement of composite fermion theory is that the strongly correlated electrons at a magnetic field (or filling factor ) turn into weakly interacting composite fermions at a magnetic field (or composite fermion filling factor ). This allows an effectively single-particle explanation of the otherwise complex many-body behavior, with the interaction between electrons manifesting as an effective kinetic energy of composite fermions. Here are some of the phenomena arising from composite fermions: [1] [2] [3]

Fermi sea

The effective magnetic field for composite fermions vanishes for , where the filling factor for electrons is . Here, composite fermions make a Fermi sea. [6] This Fermi sea has been observed at half filled Landau level in a number of experiments, which also measure the Fermi wave vector. [7] [8] [9] [10]

Cyclotron orbits

As the magnetic field is moved slightly away from , composite fermions execute semiclassical cyclotron orbits. These have been observed by coupling to surface acoustic waves, [7] resonance peaks in antidot superlattice, [8] and magnetic focusing. [9] [10] [11] The radius of the cyclotron orbits is consistent with the effective magnetic field and is sometimes an order of magnitude or more larger than the radius of the cyclotron orbit of an electron at the externally applied magnetic field . Also, the observed direction of trajectory is opposite to that of electrons when is anti-parallel to .

Cyclotron resonance

In addition to the cyclotron orbits, cyclotron resonance of composite fermions has also been observed by photoluminescence. [12]

Shubnikov de Haas oscillations

As the magnetic field is moved further away from , quantum oscillations are observed that are periodic in These are Shubnikov–de Haas oscillations of composite fermions. [13] [14] These oscillations arise from the quantization of the semiclassical cyclotron orbits of composite fermions into composite fermion Landau levels. From the analysis of the Shubnikov–de Haas experiments, one can deduce the effective mass and the quantum lifetime of composite fermions.

Integer quantum Hall effect

With further increase in or decrease in temperature and disorder, composite fermions exhibit integer quantum Hall effect. [5] The integer fillings of composite fermions, , correspond to the electrons fillings

Combined with

which are obtained by attaching vortices to holes in the lowest Landau level, these constitute the prominently observed sequences of fractions. Examples are

The fractional quantum Hall effect of electrons is thus explained as the integer quantum Hall effect of composite fermions. [5] It results in fractionally quantized Hall plateaus at

with given by above quantized values. These sequences terminate at the composite fermion Fermi sea. Note that the fractions have odd denominators, which follows from the even vorticity of composite fermions.

Fractional quantum Hall effect

The above sequences account for most, but not all, observed fractions. Other fractions have been observed, which arise from a weak residual interaction between composite fermions, and are thus more delicate. [15] A number of these are understood as fractional quantum Hall effect of composite fermions. For example, the fractional quantum Hall effect of composite fermions at produces the fraction 4/11, which does not belong to the primary sequences. [16]

Superconductivity

An even denominator fraction, has been observed. [17] Here the second Landau level is half full, but the state cannot be a Fermi sea of composite fermions, because the Fermi sea is gapless and does not show quantum Hall effect. This state is viewed as a "superconductor" of composite fermion, [18] [19] arising from a weak attractive interaction between composite fermions at this filling factor. The pairing of composite fermions opens a gap and produces a fractional quantum Hall effect.

Excitons

The neutral excitations of various fractional quantum Hall states are excitons of composite fermions, that is, particle hole pairs of composite fermions. [20] The energy dispersion of these excitons has been measured by light scattering [21] [22] and phonon scattering. [23]

Spin

At high magnetic fields the spin of composite fermions is frozen, but it is observable at relatively low magnetic fields. The fan diagram of the composite fermion Landau levels has been determined by transport, and shows both spin-up and spin-down composite fermion Landau levels. [24] The fractional quantum Hall states as well as composite fermion Fermi sea are also partially spin polarized for relatively low magnetic fields. [24] [25] [26]

Effective magnetic field

The effective magnetic field of composite fermions has been confirmed by the similarity of the fractional and the integer quantum Hall effects, observation of Fermi sea at half filled Landau level, and measurements of the cyclotron radius.

Mass

The mass of composite fermions has been determined from the measurements of: the effective cyclotron energy of composite fermions; [27] [28] the temperature dependence of Shubnikov–de Haas oscillations; [13] [14] energy of the cyclotron resonance; [12] spin polarization of the Fermi sea; [26] and quantum phase transitions between states with different spin polarizations. [24] [25] Its typical value in GaAs systems is on the order of the electron mass in vacuum. (It is unrelated to the electron band mass in GaAs, which is 0.07 of the electron mass in vacuum.)

Theoretical formulations

Much of the experimental phenomenology can be understood from the qualitative picture of composite fermions in an effective magnetic field. In addition, composite fermions also lead to a detailed and accurate microscopic theory of this quantum liquid. Two approaches have proved useful.

Trial wave functions

The following trial wave functions [5] embody the composite fermion physics:

Here is the wave function of interacting electrons at filling factor ; is the wave function for weakly interacting electrons at ; is the number of electrons or composite fermions; is the coordinate of the th particle; and is an operator that projects the wave function into the lowest Landau level. This provides an explicit mapping between the integer and the fractional quantum Hall effects. Multiplication by attaches vortices to each electron to convert it into a composite fermion. The right hand side is thus interpreted as describing composite fermions at filling factor . The above mapping gives wave functions for both the ground and excited states of the fractional quantum Hall states in terms of the corresponding known wave functions for the integral quantum Hall states. The latter do not contain any adjustable parameters for , so the FQHE wave functions do not contain any adjustable parameters at .

Comparisons with exact results show that these wave functions are quantitatively accurate. They can be used to compute a number of measurable quantities, such as the excitation gaps and exciton dispersions, the phase diagram of composite fermions with spin, the composite fermion mass, etc. For they reduce to the Laughlin wavefunction [29] at fillings .

Chern–Simons field theory

Another formulation of the composite fermion physics is through a Chern–Simons field theory, wherein flux quanta are attached to electrons by a singular gauge transformation. [6] [30] At the mean field approximation the physics of free fermions in an effective field is recovered. Perturbation theory at the level of the random phase approximation captures many of the properties of composite fermions. [31]

See also

Related Research Articles

<span class="mw-page-title-main">Quantum electrodynamics</span> Quantum field theory of electromagnetism

In particle physics, quantum electrodynamics (QED) is the relativistic quantum field theory of electrodynamics. In essence, it describes how light and matter interact and is the first theory where full agreement between quantum mechanics and special relativity is achieved. QED mathematically describes all phenomena involving electrically charged particles interacting by means of exchange of photons and represents the quantum counterpart of classical electromagnetism giving a complete account of matter and light interaction.

The quantum Hall effect is a quantized version of the Hall effect which is observed in two-dimensional electron systems subjected to low temperatures and strong magnetic fields, in which the Hall resistance Rxy exhibits steps that take on the quantized values

<span class="mw-page-title-main">Fermi liquid theory</span> Theoretical model of interacting fermions

Fermi liquid theory is a theoretical model of interacting fermions that describes the normal state of most metals at sufficiently low temperatures. The interactions among the particles of the many-body system do not need to be small. The phenomenological theory of Fermi liquids was introduced by the Soviet physicist Lev Davidovich Landau in 1956, and later developed by Alexei Abrikosov and Isaak Khalatnikov using diagrammatic perturbation theory. The theory explains why some of the properties of an interacting fermion system are very similar to those of the ideal Fermi gas, and why other properties differ.

<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 physics, an anyon is a type of quasiparticle that occurs only in two-dimensional systems, with properties much less restricted than the two kinds of standard elementary particles, fermions and bosons. In general, the operation of exchanging two identical particles, although it may cause a global phase shift, cannot affect observables. Anyons are generally classified as abelian or non-abelian. Abelian anyons play a major role in the fractional quantum Hall effect. Non-abelian anyons have not been definitively detected, although this is an active area of research.

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 . 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" Laughlin's explanation only applies to fillings where is an odd integer. The microscopic origin of the FQHE is a major research topic in condensed matter physics.

<span class="mw-page-title-main">Hofstadter's butterfly</span> Fractal describing the theorised behaviour of electrons in a magnetic field

In condensed matter physics, Hofstadter's butterfly is a graph of the spectral properties of non-interacting two-dimensional electrons in a perpendicular magnetic field in a lattice. The fractal, self-similar nature of the spectrum was discovered in the 1976 Ph.D. work of Douglas Hofstadter and is one of the early examples of modern scientific data visualization. The name reflects the fact that, as Hofstadter wrote, "the large gaps [in the graph] form a very striking pattern somewhat resembling a butterfly."

<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.

A two-dimensional electron gas (2DEG) is a scientific model in solid-state physics. It is an electron gas that is free to move in two dimensions, but tightly confined in the third. This tight confinement leads to quantized energy levels for motion in the third direction, which can then be ignored for most problems. Thus the electrons appear to be a 2D sheet embedded in a 3D world. The analogous construct of holes is called a two-dimensional hole gas (2DHG), and such systems have many useful and interesting properties.

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

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.

A topological quantum computer is a theoretical quantum computer proposed by Russian-American physicist Alexei Kitaev in 1997. It employs quasiparticles in two-dimensional systems, called anyons, whose world lines pass around one another to form braids in a three-dimensional spacetime. These braids form the logic gates that make up the computer. The advantage of a quantum computer based on quantum braids over using trapped quantum particles is that the former is much more stable. Small, cumulative perturbations can cause quantum states to decohere and introduce errors in the computation, but such small perturbations do not change the braids' topological properties. This is like the effort required to cut a string and reattach the ends to form a different braid, as opposed to a ball bumping into a wall.

<span class="mw-page-title-main">Majorana fermion</span> Fermion that is its own antiparticle

A Majorana fermion, also referred to as a Majorana particle, is a fermion that is its own antiparticle. They were hypothesised by Ettore Majorana in 1937. The term is sometimes used in opposition to a Dirac fermion, which describes fermions that are not their own antiparticles.

In solid-state physics, heavy fermion materials are a specific type of intermetallic compound, containing elements with 4f or 5f electrons in unfilled electron bands. Electrons are one type of fermion, and when they are found in such materials, they are sometimes referred to as heavy electrons. Heavy fermion materials have a low-temperature specific heat whose linear term is up to 1000 times larger than the value expected from the free electron model. The properties of the heavy fermion compounds often derive from the partly filled f-orbitals of rare-earth or actinide ions, which behave like localized magnetic moments. The name "heavy fermion" comes from the fact that the fermion behaves as if it has an effective mass greater than its rest mass. In the case of electrons, below a characteristic temperature (typically 10 K), the conduction electrons in these metallic compounds behave as if they had an effective mass up to 1000 times the free particle mass. This large effective mass is also reflected in a large contribution to the resistivity from electron-electron scattering via the Kadowaki–Woods ratio. Heavy fermion behavior has been found in a broad variety of states including metallic, superconducting, insulating and magnetic states. Characteristic examples are CeCu6, CeAl3, CeCu2Si2, YbAl3, UBe13 and UPt3.

The quantum spin Hall state is a state of matter proposed to exist in special, two-dimensional semiconductors that have a quantized spin-Hall conductance and a vanishing charge-Hall conductance. The quantum spin Hall state of matter is the cousin of the integer quantum Hall state, and that does not require the application of a large magnetic field. The quantum spin Hall state does not break charge conservation symmetry and spin- conservation symmetry.

In particle physics and string theory (M-theory), the ADD model, also known as the model with large extra dimensions (LED), is a model framework that attempts to solve the hierarchy problem. The model tries to explain this problem by postulating that our universe, with its four dimensions, exists on a membrane in a higher dimensional space. It is then suggested that the other forces of nature operate within this membrane and its four dimensions, while gravitons can propagate across the extra dimensions. This would explain why gravity is very weak compared to the other fundamental forces. The size of the dimensions in ADD is around the order of the TeV scale, which results in it being experimentally probeable by current colliders, unlike many exotic extra dimensional hypotheses that have the relevant size around the Planck scale.

<span class="mw-page-title-main">Luttinger's theorem</span>

In condensed matter physics, Luttinger's theorem is a result derived by J. M. Luttinger and J. C. Ward in 1960 that has broad implications in the field of electron transport. It arises frequently in theoretical models of correlated electrons, such as the high-temperature superconductors, and in photoemission, where a metal's Fermi surface can be directly observed.

<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.

James (Jim) P. Eisenstein is the Frank J. Roshek Professor of Physics and Applied Physics at the physics department of California Institute of Technology.

<span class="mw-page-title-main">Electronic properties of graphene</span>

Graphene is a semimetal whose conduction and valence bands meet at the Dirac points, which are six locations in momentum space, the vertices of its hexagonal Brillouin zone, divided into two non-equivalent sets of three points. The two sets are labeled K and K'. The sets give graphene a valley degeneracy of gv = 2. By contrast, for traditional semiconductors the primary point of interest is generally Γ, where momentum is zero. Four electronic properties separate it from other condensed matter systems.

The term Dirac matter refers to a class of condensed matter systems which can be effectively described by the Dirac equation. Even though the Dirac equation itself was formulated for fermions, the quasi-particles present within Dirac matter can be of any statistics. As a consequence, Dirac matter can be distinguished in fermionic, bosonic or anyonic Dirac matter. Prominent examples of Dirac matter are Graphene, topological insulators, Dirac semimetals, Weyl semimetals, various high-temperature superconductors with -wave pairing and liquid Helium-3. The effective theory of such systems is classified by a specific choice of the Dirac mass, the Dirac velocity, the Dirac matrices and the space-time curvature. The universal treatment of the class of Dirac matter in terms of an effective theory leads to a common features with respect to the density of states, the heat capacity and impurity scattering.

References

  1. 1 2 3 J.K. Jain (2007). Composite Fermions. New York: Cambridge University Press. ISBN   978-0-521-86232-5.
  2. 1 2 3 O. Heinonen, ed. (1998). Composite Fermions. Singapore: World Scientific. ISBN   978-981-02-3592-5.
  3. 1 2 3 S. Das Sarma; A. Pinczuk, eds. (1996). Perspectives in Quantum Hall Effects: Novel Quantum Liquids in Low Dimensional Semiconductor Structures. New York: Wiley-VCH. ISBN   978-0-471-11216-7.
  4. D.C. Tsui; H.L. Stormer; A.C. Gossard (1982). "Two-dimensional magnetotransport in the extreme quantum limit". Physical Review Letters . 48 (22): 1559. Bibcode:1982PhRvL..48.1559T. doi: 10.1103/PhysRevLett.48.1559 .
  5. 1 2 3 4 J.K. Jain (1989). "Composite fermion approach for fractional quantum Hall effect". Physical Review Letters . 63 (2): 199–202. Bibcode:1989PhRvL..63..199J. doi:10.1103/PhysRevLett.63.199. PMID   10040805.
  6. 1 2 B.I. Halperin; P.A. Lee; N. Read (1993). "Theory of the half-filled Landau level". Physical Review B . 47 (12): 7312–7343. arXiv: cond-mat/9501090 . Bibcode:1993PhRvB..47.7312H. doi:10.1103/PhysRevB.47.7312. PMID   10004728.
  7. 1 2 R.L. Willett; R.R. Ruel; K.W. West; L.N. Pfeiffer (1993). "Experimental demonstration of a Fermi surface at one-half filling of the lowest Landau level". Physical Review Letters . 71 (23): 3846–3849. Bibcode:1993PhRvL..71.3846W. doi:10.1103/PhysRevLett.71.3846. PMID   10055088.
  8. 1 2 W. Kang; H.L. Stormer; L.N. Pfeiffer; K.W. Baldwin; K.W. West (1993). "How Real are composite fermions?". Physical Review Letters . 71 (23): 3850–3853. Bibcode:1993PhRvL..71.3850K. doi:10.1103/PhysRevLett.71.3850. PMID   10055089.
  9. 1 2 V.J. Goldman; B. Su; J.K. Jain (1994). "Detection of composite fermions by magnetic focusing". Physical Review Letters . 72 (13): 2065–2068. Bibcode:1994PhRvL..72.2065G. doi:10.1103/PhysRevLett.72.2065. PMID   10055779.
  10. 1 2 J.H. Smet; D. Weiss; R.H. Blick; G. Lütjering; K. von Klitzing; R. Fleischmann; R. Ketzmerick; T. Geisel; G. Weimann (1996). "Magnetic focusing of composite fermions through arrays of cavities". Physical Review Letters . 77 (11): 2272–2275. Bibcode:1996PhRvL..77.2272S. doi:10.1103/PhysRevLett.77.2272. PMID   10061902. S2CID   20584064.
  11. J. H. Smet; S. Jobst; K. von Klitzing; D. Weiss; W. Wegscheider; V. Umansky (1999). "Commensurate composite fermions in weak periodic electrostatic potentials: Direct evidence of a periodic effective magnetic field" (PDF). Physical Review Letters . 83 (13): 2620. Bibcode:1999PhRvL..83.2620S. doi:10.1103/PhysRevLett.83.2620. S2CID   122014617.
  12. 1 2 I.V. Kukushkin; J.H. Smet; D. Schuh; W. Wegscheider; K. von Klitzing (2007). "Dispersion of the composite-fermion cyclotron resonance mode". Physical Review Letters . 98 (6): 066403. Bibcode:2007PhRvL..98f6403K. doi:10.1103/PhysRevLett.98.066403. PMID   17358964.
  13. 1 2 D.R. Leadley; R.J. Nicholas; C.T. Foxon; J.J. Harris (1994). "Measurement of the effective mass and scattering times of composite fermions from magnetotransport analysis". Physical Review Letters . 72 (12): 1906–1909. Bibcode:1994PhRvL..72.1906L. doi:10.1103/PhysRevLett.72.1906. PMID   10055734.
  14. 1 2 R.R. Du; H.L. Stormer; D.C. Tsui; L.N. Pfeiffer; K.W. West (1994). "Shubnikov–de Haas oscillations around Landaulevel filling". Solid State Communications. 90 (2): 71. Bibcode:1994SSCom..90...71D. doi:10.1016/0038-1098(94)90934-2.
  15. W. Pan; H.L. Stormer; D.C. Tsui; L.N. Pfeiffer; K.W. Baldwin; K.W. West (2003). "Fractional quantum Hall effect of composite fermions". Physical Review Letters . 90 (1): 016801. arXiv: cond-mat/0303429 . Bibcode:2003PhRvL..90a6801P. doi:10.1103/PhysRevLett.90.016801. PMID   12570639. S2CID   2265408.
  16. C.-C. Chang; J.K. Jain (2004). "Microscopic origin of the next generation fractional quantum Hall effect". Physical Review Letters . 92 (19): 196806. arXiv: cond-mat/0404079 . Bibcode:2004PhRvL..92s6806C. doi:10.1103/PhysRevLett.92.196806. PMID   15169434. S2CID   20862603.
  17. R. Willett; J.P. Eisenstein; H.L. Stormer; D.C. Tsui; A.C. Gossard; J.H. England (1987). "Observation of an even-denominator quantum number in the fractional quantum Hall effect" (PDF). Physical Review Letters . 59 (15): 1776–1779. Bibcode:1987PhRvL..59.1776W. doi:10.1103/PhysRevLett.59.1776. PMID   10035326.
  18. G. Moore; N. Read (1991). "Nonabelions in the fractional quantum Hall effect" (PDF). Nuclear Physics B. 360 (2): 362. Bibcode:1991NuPhB.360..362M. doi:10.1016/0550-3213(91)90407-O.
  19. N. Read; D. Green (2000). "Paired states of fermions in two dimensions with breaking of parity and time reversal symmetries and the fractional quantum Hall effect". Physical Review B . 61 (15): 10267. arXiv: cond-mat/9906453 . Bibcode:2000PhRvB..6110267R. doi:10.1103/PhysRevB.61.10267. S2CID   119427877.
  20. V.W. Scarola; K. Park; J.K. Jain (2000). "Rotons of composite fermions: Comparison between theory and experiment". Physical Review B . 61 (19): 13064. arXiv: cond-mat/9910491 . Bibcode:2000PhRvB..6113064S. doi:10.1103/PhysRevB.61.13064.
  21. M. Kang; A. Pinczuk; B.S. Dennis; L.N. Pfeiffer; K.W. West (2001). "Observation of multiple magnetorotons in the fractional quantum Hall effect". Physical Review Letters . 86 (12): 2637–40. Bibcode:2001PhRvL..86.2637K. doi:10.1103/PhysRevLett.86.2637. PMID   11289999.
  22. I. Dujovne; A. Pinczuk; M. Kang; B.S. Dennis; L.N. Pfeiffer; K.W. West (2005). "Composite-fermion spin excitations at approaches ½: Interactions in the Fermi sea". Physical Review Letters . 95 (5): 056808. Bibcode:2005PhRvL..95e6808D. doi:10.1103/PhysRevLett.95.056808. PMID   16090907.
  23. F. Schulze-Wischeler; F. Hohls; U. Zeitler; D. Reuter; A.D. Wieck; R.J. Haug (2004). "Phonon excitations of composite fermion Landau levels". Physical Review Letters . 93 (2): 026801. arXiv: cond-mat/0403072 . Bibcode:2004PhRvL..93b6801S. doi:10.1103/PhysRevLett.93.026801. PMID   15323936.
  24. 1 2 3 R.R. Du; A.S. Yeh; H.L. Stormer; D.C. Tsui; L.N. Pfeiffer; K.W. West (1995). "Fractional quantum Hall effect around : Composite fermions with a spin". Physical Review Letters . 75 (21): 3926–3929. Bibcode:1995PhRvL..75.3926D. doi:10.1103/PhysRevLett.75.3926. PMID   10059766.
  25. 1 2 I.V. Kukushkin; K. von Klitzing; K. Eberl (1999). "Spin polarization of composite fermions: Measurements of the Fermi energy". Physical Review Letters . 82 (18): 3665. Bibcode:1999PhRvL..82.3665K. doi:10.1103/PhysRevLett.82.3665.
  26. 1 2 S. Melinte; N. Freytag; M. Horvatic; C. Berthier; L.P. Levy; V. Bayot; M. Shayegan (2000). "NMR determination of 2D electron spin polarization at ". Physical Review Letters . 84 (2): 354–7. arXiv: cond-mat/9908098 . Bibcode:2000PhRvL..84..354M. doi:10.1103/PhysRevLett.84.354. PMID   11015909. S2CID   42918257.
  27. R.R. Du; H.L. Stormer; D.C. Tsui; L.N. Pfeiffer; K.W. Baldwin; K.W. West (1993). "Experimental evidence for new particles in the fractional quantum Hall effect". Physical Review Letters . 70 (19): 2944–2947. Bibcode:1993PhRvL..70.2944D. doi:10.1103/PhysRevLett.70.2944. PMID   10053693.
  28. H.C. Manoharan; M. Shayegan; S.J. Klepper (1994). "Signatures of a novel Fermi liquid in a two-dimensional composite particle model". Physical Review Letters . 73 (24): 3270–3273. Bibcode:1994PhRvL..73.3270M. doi:10.1103/PhysRevLett.73.3270. PMID   10057334.
  29. R.B. Laughlin (1983). "Anomalous Quantum Hall Effect: An Incompressible Quantum Fluid with Fractionally Charged Excitations". Physical Review Letters . 50 (18): 1395. Bibcode:1983PhRvL..50.1395L. doi:10.1103/PhysRevLett.50.1395. S2CID   120080343.
  30. A. Lopez; E. Fradkin (1991). "Fractional quantum Hall effect and Chern–Simons gauge theories". Physical Review B . 44 (10): 5246–5262. Bibcode:1991PhRvB..44.5246L. doi:10.1103/PhysRevB.44.5246. PMID   9998334.
  31. S.H. Simon; B.I. Halperin (1993). "Finite-wave-vector electromagnetic response of fractional quantized Hall states". Physical Review B . 48 (23): 17368–17387. arXiv: cond-mat/9307048 . Bibcode:1993PhRvB..4817368S. doi:10.1103/PhysRevB.48.17368. PMID   10008349. S2CID   32195345.