Anderson localization

In condensed matter physics, Anderson localization (also known as strong localization)[1] is the absence of diffusion of waves in a disordered medium. This phenomenon is named after the American physicist P. W. Anderson, who was the first to suggest that electron localization is possible in a lattice potential, provided that the degree of randomness (disorder) in the lattice is sufficiently large, as can be realized for example in a semiconductor with impurities or defects.[2]

Anderson localization is a general wave phenomenon that applies to the transport of electromagnetic waves, acoustic waves, quantum waves, spin waves, etc. This phenomenon is to be distinguished from weak localization, which is the precursor effect of Anderson localization (see below), and from Mott localization, named after Sir Nevill Mott, where the transition from metallic to insulating behaviour is not due to disorder, but to a strong mutual Coulomb repulsion of electrons. It is shown that strong disorder can be employed to obtain high-quality wavefront due to the Anderson localization phenomenon in a transverse Anderson localizing optical fiber.[3]

Introduction

In the original Anderson tight-binding model, the evolution of the wave function ψ on the d-dimensional lattice Zd is given by the Schrödinger equation

where the Hamiltonian H is given by

with Ej random and independent, and potential V(r) falling off as r−2 at infinity. For example, one may take Ej uniformly distributed in [W,   +W], and

Starting with ψ0 localised at the origin, one is interested in how fast the probability distribution diffuses. Anderson's analysis shows the following:

  • if d is 1 or 2 and W is arbitrary, or if d 3 and W/ħ is sufficiently large, then the probability distribution remains localized:
uniformly in t. This phenomenon is called Anderson localization.
  • if d 3 and W/ħ is small,
where D is the diffusion constant.

Analysis

Example of a multifractal electronic eigenstate at the Anderson localization transition in a system with 1367631 atoms.

The phenomenon of Anderson localization, particularly that of weak localization, finds its origin in the wave interference between multiple-scattering paths. In the strong scattering limit, the severe interferences can completely halt the waves inside the disordered medium.

For non-interacting electrons, a highly successful approach was put forward in 1979 by Abrahams et al.[4] This scaling hypothesis of localization suggests that a disorder-induced metal-insulator transition (MIT) exists for non-interacting electrons in three dimensions (3D) at zero magnetic field and in the absence of spin-orbit coupling. Much further work has subsequently supported these scaling arguments both analytically and numerically (Brandes et al., 2003; see Further Reading). In 1D and 2D, the same hypothesis shows that there are no extended states and thus no MIT. However, since 2 is the lower critical dimension of the localization problem, the 2D case is in a sense close to 3D: states are only marginally localized for weak disorder and a small spin-orbit coupling can lead to the existence of extended states and thus an MIT. Consequently, the localization lengths of a 2D system with potential-disorder can be quite large so that in numerical approaches one can always find a localization-delocalization transition when either decreasing system size for fixed disorder or increasing disorder for fixed system size.

Most numerical approaches to the localization problem use the standard tight-binding Anderson Hamiltonian with onsite-potential disorder. Characteristics of the electronic eigenstates are then investigated by studies of participation numbers obtained by exact diagonalization, multifractal properties, level statistics and many others. Especially fruitful is the transfer-matrix method (TMM) which allows a direct computation of the localization lengths and further validates the scaling hypothesis by a numerical proof of the existence of a one-parameter scaling function. Direct numerical solution of Maxwell equations to demonstrate Anderson localization of light has been implemented (Conti and Fratalocchi, 2008).

Experimental evidence

Two reports of Anderson localization of light in 3D random media exist up to date (Wiersma et al., 1997 and Storzer et al., 2006; see Further Reading), even though absorption complicates interpretation of experimental results (Scheffold et al., 1999). Anderson localization can also be observed in a perturbed periodic potential where the transverse localization of light is caused by random fluctuations on a photonic lattice. Experimental realizations of transverse localization were reported for a 2D lattice (Schwartz et al., 2007) and a 1D lattice (Lahini et al., 2006). Transverse Anderson localization of light has also been demonstrated in an optical fiber medium (Karbasi et al., 2012) and a biological medium (Choi et al., 2018), and has also been used to transport images through the fiber (Karbasi et al., 2014). It has also been observed by localization of a Bose–Einstein condensate in a 1D disordered optical potential (Billy et al., 2008; Roati et al., 2008). Anderson localization of elastic waves in a 3D disordered medium has been reported (Hu et al., 2008). The observation of the MIT has been reported in a 3D model with atomic matter waves (Chabé et al., 2008). The MIT, associated with the nonpropagative electron waves has been reported in a cm sized crystal (Ying et al., 2016) Random lasers can operate using this phenomenon.

Comparison with diffusion

Standard diffusion has no localization property, being in disagreement with quantum predictions. However, it turns out that it is based on approximation of the principle of maximum entropy, which says that the probability distribution which best represents the current state of knowledge is the one with largest entropy. This approximation is repaired in Maximal Entropy Random Walk, also repairing the disagreement: it turns out to lead to exactly the quantum ground state stationary probability distribution with its strong localization properties.[5][6]

Notes

  1. Fabian Teichert, Andreas Zienert, Jörg Schuster, Michael Schreiber (2014). "Strong localization in defective carbon nanotubes: a recursive Green's function study". New Journal of Physics. 16 (12): 123026. arXiv:1705.01757. Bibcode:2014NJPh...16l3026T. doi:10.1088/1367-2630/16/12/123026.
  2. Anderson, P. W. (1958). "Absence of Diffusion in Certain Random Lattices". Phys. Rev. 109 (5): 1492&ndash, 1505. Bibcode:1958PhRv..109.1492A. doi:10.1103/PhysRev.109.1492.
  3. Abaie, Behnam; Peysokhan, Mostafa; Zhao, Jian; Antonio-Lopez, Jose E.; Amezcua-Correa, Rodrigo; Schülzgen, Axel; Mafi, Arash (2018-08-20). "Disorder-induced high-quality wavefront in an Anderson localizing optical fiber". Optica. 5 (8): 984–987. doi:10.1364/OPTICA.5.000984. ISSN 2334-2536.
  4. Abrahams, E.; Anderson, P.W.; Licciardello, D.C.; Ramakrishnan, T.V. (1979). "Scaling Theory of Localization: Absence of Quantum Diffusion in Two Dimensions". Phys. Rev. Lett. 42 (10): 673&ndash, 676. Bibcode:1979PhRvL..42..673A. doi:10.1103/PhysRevLett.42.673.
  5. Z. Burda, J. Duda, J. M. Luck, and B. Waclaw, Localization of the Maximal Entropy Random Walk, Phys. Rev. Lett., 2009.
  6. J. Duda, Extended Maximal Entropy Random Walk, PhD Thesis, 2012.

Further reading

  • Brandes, T. & Kettemann, S. (2003). "The Anderson Transition and its Ramifications --- Localisation, Quantum Interference, and Interactions". Berlin: Springer Verlag
  • Wiersma, Diederik S.; et al. (1997). "Localization of light in a disordered medium". Nature. 390 (6661): 671&ndash, 673. Bibcode:1997Natur.390..671W. doi:10.1038/37757.
  • Störzer, Martin; et al. (2006). "Observation of the critical regime near Anderson localization of light". Phys. Rev. Lett. 96 (6): 063904. arXiv:cond-mat/0511284. Bibcode:2006PhRvL..96f3904S. doi:10.1103/PhysRevLett.96.063904. PMID 16605998.
  • Scheffold, Frank; et al. (1999). "Localization or classical diffusion of light?". Nature. 398 (6724): 206&ndash, 207. Bibcode:1999Natur.398..206S. doi:10.1038/18347.
  • Schwartz, T.; et al. (2007). "Transport and Anderson Localization in disordered two-dimensional Photonic Lattices". Nature. 446 (7131): 52&ndash, 55. Bibcode:2007Natur.446...52S. doi:10.1038/nature05623. PMID 17330037.
  • Lahini, Y.; et al. (2008). "Anderson localization and nonlinearity in one dimensional disordered photonic lattices". Physical Review Letters. 100 (1): 013906. arXiv:0704.3788. Bibcode:2008PhRvL.100a3906L. doi:10.1103/PhysRevLett.100.013906. PMID 18232768.
  • Karbasi, S.; et al. (2012). "Observation of transverse Anderson localization in an optical fiber". Optics Letters. 37 (12): 2304–6. Bibcode:2012OptL...37.2304K. doi:10.1364/OL.37.002304. PMID 22739889.
  • Karbasi, S.; et al. (2014). "Image transport through a disordered optical fibre mediated by transverse Anderson localization". Nature Communications. 5: 3362. arXiv:1307.4160. Bibcode:2014NatCo...5E3362K. doi:10.1038/ncomms4362. PMID 24566557.
  • Billy, Juliette; et al. (2008). "Direct observation of Anderson localization of matter waves in a controlled disorder". Nature. 453 (7197): 891&ndash, 894. arXiv:0804.1621. Bibcode:2008Natur.453..891B. doi:10.1038/nature07000. PMID 18548065.
  • Roati, Giacomo; et al. (2008). "Anderson localization of a non-interacting Bose-Einstein condensate". Nature. 453 (7197): 895&ndash, 898. arXiv:0804.2609. Bibcode:2008Natur.453..895R. doi:10.1038/nature07071. PMID 18548066.
  • Ludlam, J. J.; et al. (2005). "Universal features of localized eigenstates in disordered systems". Journal of Physics: Condensed Matter. 17 (30): L321–L327. Bibcode:2005JPCM...17L.321L. doi:10.1088/0953-8984/17/30/L01.
  • Conti, C; A. Fratalocchi (2008). "Dynamic light diffusion, three-dimensional Anderson localization and lasing in inverted opals". Nature Physics. 4 (10): 794&ndash, 798. arXiv:0802.3775. Bibcode:2008NatPh...4..794C. doi:10.1038/nphys1035.
  • Hu, Hefei; et al. (2008). "Localization of ultrasound in a three-dimensional elastic network". Nature Physics. 4 (12): 945–948. arXiv:0805.1502. Bibcode:2008NatPh...4..945H. doi:10.1038/nphys1101.
  • Chabé, J.; et al. (2008). "Experimental Observation of the Anderson Metal-Insulator Transition with Atomic Matter Waves". Phys. Rev. Lett. 101 (25): 255702. arXiv:0709.4320. Bibcode:2008PhRvL.101y5702C. doi:10.1103/PhysRevLett.101.255702. PMID 19113725.
  • Ying, Tianping; et al. (2016). "Anderson localization of electrons in single crystals: LixFe7Se8". Science Advances. 2 (2): e1501283. Bibcode:2016SciA....2E1283Y. doi:10.1126/sciadv.1501283. PMC 4788481. PMID 26989781.
  • Choi, Seung Ho; et al. (2018). "Anderson light localization in biological nanostructures of native silk". Nature Communications. 9 (1): 452. doi:10.1038/s41467-017-02500-5. PMC 5792459. PMID 29386508.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.