fbpx
Wikipedia

Bose–Hubbard model

The Bose–Hubbard model gives a description of the physics of interacting spinless bosons on a lattice. It is closely related to the Hubbard model that originated in solid-state physics as an approximate description of superconducting systems and the motion of electrons between the atoms of a crystalline solid. The model was introduced by Gersch and Knollman[1] in 1963 in the context of granular superconductors. (The term 'Bose' in its name refers to the fact that the particles in the system are bosonic.) The model rose to prominence in the 1980s after it was found to capture the essence of the superfluid-insulator transition in a way that was much more mathematically tractable than fermionic metal-insulator models.[2][3][4]

The Bose–Hubbard model can be used to describe physical systems such as bosonic atoms in an optical lattice,[5] as well as certain magnetic insulators.[6][7] Furthermore, it can be generalized and applied to Bose–Fermi mixtures, in which case the corresponding Hamiltonian is called the Bose–Fermi–Hubbard Hamiltonian.

Hamiltonian edit

The physics of this model is given by the Bose–Hubbard Hamiltonian:

 .

Here,   denotes summation over all neighboring lattice sites   and  , while   and   are bosonic creation and annihilation operators such that   gives the number of particles on site  . The model is parametrized by the hopping amplitude   that describes boson mobility in the lattice, the on-site interaction   which can be attractive ( ) or repulsive ( ), and the chemical potential  , which essentially sets the number of particles. If unspecified, typically the phrase 'Bose–Hubbard model' refers to the case where the on-site interaction is repulsive.

This Hamiltonian has a global   symmetry, which means that it is invariant (its physical properties are unchanged) by the transformation  . In a superfluid phase, this symmetry is spontaneously broken.

Hilbert space edit

The dimension of the Hilbert space of the Bose–Hubbard model is given by  , where   is the total number of particles, while   denotes the total number of lattice sites. At fixed   or  , the Hilbert space dimension   grows polynomially, but at a fixed density of   bosons per site, it grows exponentially as  . Analogous Hamiltonians may be formulated to describe spinless fermions (the Fermi-Hubbard model) or mixtures of different atom species (Bose–Fermi mixtures, for example). In the case of a mixture, the Hilbert space is simply the tensor product of the Hilbert spaces of the individual species. Typically additional terms are included to model interaction between species.

Phase diagram edit

At zero temperature, the Bose–Hubbard model (in the absence of disorder) is in either a Mott insulating state at small  , or in a superfluid state at large  .[8] The Mott insulating phases are characterized by integer boson densities, by the existence of an energy gap for particle-hole excitations, and by zero compressibility. The superfluid is characterized by long-range phase coherence, a spontaneous breaking of the Hamiltonian's continuous   symmetry, a non-zero compressibility and superfluid susceptibility. At non-zero temperature, in certain parameter regimes a regular fluid phase appears that does not break the   symmetry and does not display phase coherence. Both of these phases have been experimentally observed in ultracold atomic gases.[9]

In the presence of disorder, a third, "Bose glass" phase exists.[4] The Bose glass is a Griffiths phase, and can be thought of as a Mott insulator containing rare 'puddles' of superfluid. These superfluid pools are not interconnected, so the system remains insulating, but their presence significantly changes model thermodynamics. The Bose glass phase is characterized by finite compressibility, the absence of a gap, and by an infinite superfluid susceptibility.[4] It is insulating despite the absence of a gap, as low tunneling prevents the generation of excitations which, although close in energy, are spatially separated. The Bose glass has a non-zero Edwards–Anderson order parameter[10][11] and has been suggested (but not proven) to display replica symmetry breaking.[12]

Mean-field theory edit

The phases of the clean Bose–Hubbard model can be described using a mean-field Hamiltonian:[13]

 
where   is the lattice co-ordination number. This can be obtained from the full Bose–Hubbard Hamiltonian by setting   where  , neglecting terms quadratic in   (assumedly infinitesimal) and relabelling  . Because this decoupling breaks the   symmetry of the initial Hamiltonian for all non-zero values of  , this parameter acts as a superfluid order parameter. For simplicity, this decoupling assumes   to be the same on every site, which precludes exotic phases such as supersolids or other inhomogeneous phases. (Other decouplings are possible.) The phase diagram can be determined by calculating the energy of this mean-field Hamiltonian using second-order perturbation theory and finding the condition for which  . To do this, the Hamiltonian is written as a site-local piece plus a perturbation:
 
where the bilinear terms   and its conjugate are treated as the perturbation. The order parameter   is assumed to be small near the phase transition. The local term is diagonal in the Fock basis, giving the zeroth-order energy contribution:
 
where   is an integer that labels the filling of the Fock state. The perturbative piece can be treated with second-order perturbation theory, which leads to:
 
The energy can be expressed as a series expansion in even powers of the order parameter (also known as the Landau formalism):
 
After doing so, the condition for the mean-field, second-order phase transition between the Mott insulator and the superfluid phase is given by:
 
where the integer   describes the filling of the   Mott insulating lobe. Plotting the line   for different integer values of   generates the boundary of the different Mott lobes, as shown in the phase diagram.[4]

Implementation in optical lattices edit

Ultracold atoms in optical lattices are considered a standard realization of the Bose–Hubbard model. The ability to tune model parameters using simple experimental techniques and the lack of the lattice dynamics that are present in solid-state electronic systems mean that ultracold atoms offer a clean, controllable realisation of the Bose–Hubbard model.[14][5] The biggest downside with optical lattice technology is the trap lifetime, with atoms typically trapped for only a few tens of seconds.

To see why ultracold atoms offer such a convenient realization of Bose–Hubbard physics, the Bose–Hubbard Hamiltonian can be derived starting from the second quantized Hamiltonian that describes a gas of ultracold atoms in the optical lattice potential. This Hamiltonian is given by:

 ,

where   is the optical lattice potential,   is the (contact) interaction amplitude, and   is the chemical potential. The tight binding approximation results in the substitution  , which leads to the Bose–Hubbard Hamiltonian the physics are restricted to the lowest band ( ) and the interactions are local at the level of the discrete mode. Mathematically, this can be stated as the requirement that   except for case  . Here,   is a Wannier function for a particle in an optical lattice potential localized around site   of the lattice and for the  th Bloch band.[15]

Subtleties and approximations edit

The tight-binding approximation significantly simplifies the second quantized Hamiltonian, though it introduces several limitations at the same time:

  • For single-site states with several particles in a single state, the interactions may couple to higher Bloch bands, which contradicts base assumptions. Still, a single band model is able to address low-energy physics of such a setting but with parameters U and J becoming density-dependent. Instead of one parameter U, the interaction energy of n particles may be described by   close, but not equal to U.[15]
  • When considering (fast) lattice dynamics, additional terms are added to the Hamiltonian so that the time-dependent Schrödinger equation is obeyed in the (time-dependent) Wannier function basis. The terms come from the Wannier functions' time dependence.[16][17] Otherwise, the lattice dynamics may be incorporated by making the key parameters of the model time-dependent, varying with the instantaneous value of the optical potential.

Experimental results edit

Quantum phase transitions in the Bose–Hubbard model were experimentally observed by Greiner et al.,[9] and density dependent interaction parameters   were observed by Immanuel Bloch's group.[18] Single-atom resolution imaging of the Bose–Hubbard model has been possible since 2009 using quantum gas microscopes.[19][20][21]

Further applications edit

The Bose–Hubbard model is of interest in the field of quantum computation and quantum information. Entanglement of ultra-cold atoms can be studied using this model.[22]

Numerical simulation edit

In the calculation of low energy states the term proportional to   means that large occupation of a single site is improbable, allowing for truncation of local Hilbert space to states containing at most   particles. Then the local Hilbert space dimension is   The dimension of the full Hilbert space grows exponentially with the number of lattice sites, limiting exact computer simulations of the entire Hilbert space to systems of 15-20 particles in 15-20 lattice sites.[citation needed] Experimental systems contain several million sites, with average filling above unity.[citation needed]

One-dimensional lattices may be studied using density matrix renormalization group (DMRG) and related techniques such as time-evolving block decimation (TEBD). This includes calculating the ground state of the Hamiltonian for systems of thousands of particles on thousands of lattice sites, and simulating its dynamics governed by the time-dependent Schrödinger equation. Recently,[when?] two dimensional lattices have been studied using projected entangled pair states, a generalization of matrix product states in higher dimensions, both for the ground state[23] and finite temperature.[24]

Higher dimensions are significantly more difficult due to the rapid growth of entanglement.[25]

All dimensions may be treated by quantum Monte Carlo algorithms,[citation needed] which provide a way to study properties of the Hamiltonian's thermal states, and in particular the ground state.

Generalizations edit

Bose–Hubbard-like Hamiltonians may be derived for different physical systems containing ultracold atom gas in the periodic potential. They include:

  • systems with longer-ranged density-density interactions of the form  , which may stabilise a supersolid phase for certain parameter values
  • dimerised magnets, where spin-1/2 electrons are bound together in pairs called dimers that have bosonic excitation statistics and are described by a Bose–Hubbard model
  • long-range dipolar interaction[26]
  • systems with interaction-induced tunneling terms  [27]
  • internal spin structure of atoms, for example due to trapping an entire degenerate manifold of hyperfine spin states (for F=1 it leads to the spin-1 Bose–Hubbard model)[28][clarification needed]
  • situations where the gas experiences an additional potential—for example, in disordered systems.[29] The disorder might be realised by a speckle pattern, or using a second, incommensurate, weaker, optical lattice. In the latter case inclusion of the disorder amounts to including extra term of the form:  .

See also edit

References edit

  1. ^ Gersch, H.; Knollman, G. (1963). "Quantum Cell Model for Bosons". Physical Review. 129 (2): 959. Bibcode:1963PhRv..129..959G. doi:10.1103/PhysRev.129.959.
  2. ^ Ma, M.; Halperin, B. I.; Lee, P. A. (1986-09-01). "Strongly disordered superfluids: Quantum fluctuations and critical behavior". Physical Review B. 34 (5): 3136–3143. Bibcode:1986PhRvB..34.3136M. doi:10.1103/PhysRevB.34.3136. PMID 9940047.
  3. ^ Giamarchi, T.; Schulz, H. J. (1988-01-01). "Anderson localization and interactions in one-dimensional metals". Physical Review B. 37 (1): 325–340. Bibcode:1988PhRvB..37..325G. doi:10.1103/PhysRevB.37.325. PMID 9943580.
  4. ^ a b c d Fisher, Matthew P. A.; Grinstein, G.; Fisher, Daniel S. (1989). "Boson localization and the superfluid-insulator transition" (PDF). Physical Review B. 40 (1): 546–70. Bibcode:1989PhRvB..40..546F. doi:10.1103/PhysRevB.40.546. PMID 9990946.,
  5. ^ a b Jaksch, D.; Zoller, P. (2005). "The cold atom Hubbard toolbox". Annals of Physics. 315 (1): 52. arXiv:cond-mat/0410614. Bibcode:2005AnPhy.315...52J. CiteSeerX 10.1.1.305.9031. doi:10.1016/j.aop.2004.09.010. S2CID 12352119.
  6. ^ Giamarchi, Thierry; Rüegg, Christian; Tchernyshyov, Oleg (2008). "Bose–Einstein condensation in magnetic insulators". Nature Physics. 4 (3): 198–204. arXiv:0712.2250. Bibcode:2008NatPh...4..198G. doi:10.1038/nphys893. S2CID 118661914.
  7. ^ Zapf, Vivien; Jaime, Marcelo; Batista, C. D. (2014-05-15). "Bose-Einstein condensation in quantum magnets". Reviews of Modern Physics. 86 (2): 563–614. Bibcode:2014RvMP...86..563Z. doi:10.1103/RevModPhys.86.563.
  8. ^ Kühner, T.; Monien, H. (1998). "Phases of the one-dimensional Bose-Hubbard model". Physical Review B. 58 (22): R14741. arXiv:cond-mat/9712307. Bibcode:1998PhRvB..5814741K. doi:10.1103/PhysRevB.58.R14741. S2CID 118911555.
  9. ^ a b Greiner, Markus; Mandel, Olaf; Esslinger, Tilman; Hänsch, Theodor W.; Bloch, Immanuel (2002). "Quantum phase transition from a superfluid to a Mott insulator in a gas of ultracold atoms". Nature. 415 (6867): 39–44. Bibcode:2002Natur.415...39G. doi:10.1038/415039a. PMID 11780110. S2CID 4411344.
  10. ^ Morrison, S.; Kantian, A.; Daley, A. J.; Katzgraber, H. G.; Lewenstein, M.; Büchler, H. P.; Zoller, P. (2008). "Physical replicas and the Bose glass in cold atomic gases". New Journal of Physics. 10 (7): 073032. arXiv:0805.0488. Bibcode:2008NJPh...10g3032M. doi:10.1088/1367-2630/10/7/073032. ISSN 1367-2630. S2CID 2822103.
  11. ^ Thomson, S. J.; Walker, L. S.; Harte, T. L.; Bruce, G. D. (2016-11-03). "Measuring the Edwards-Anderson order parameter of the Bose glass: A quantum gas microscope approach". Physical Review A. 94 (5): 051601. arXiv:1607.05254. Bibcode:2016PhRvA..94e1601T. doi:10.1103/PhysRevA.94.051601. S2CID 56027819.
  12. ^ Thomson, S. J.; Krüger, F. (2014). "Replica symmetry breaking in the Bose glass". EPL. 108 (3): 30002. arXiv:1312.0515. Bibcode:2014EL....10830002T. doi:10.1209/0295-5075/108/30002. S2CID 56307253.
  13. ^ Sachdev, Subir (2011). Quantum phase transitions. Cambridge University Press. ISBN 9780521514682. OCLC 693207153.
  14. ^ Jaksch, D.; Bruder, C.; Cirac, J.; Gardiner, C.; Zoller, P. (1998). "Cold Bosonic Atoms in Optical Lattices". Physical Review Letters. 81 (15): 3108. arXiv:cond-mat/9805329. Bibcode:1998PhRvL..81.3108J. doi:10.1103/PhysRevLett.81.3108. S2CID 55578669.
  15. ^ a b Lühmann, D. S. R.; Jürgensen, O.; Sengstock, K. (2012). "Multi-orbital and density-induced tunneling of bosons in optical lattices". New Journal of Physics. 14 (3): 033021. arXiv:1108.3013. Bibcode:2012NJPh...14c3021L. doi:10.1088/1367-2630/14/3/033021. S2CID 119216031.
  16. ^ Sakmann, K.; Streltsov, A. I.; Alon, O. E.; Cederbaum, L. S. (2011). "Optimal time-dependent lattice models for nonequilibrium dynamics". New Journal of Physics. 13 (4): 043003. arXiv:1006.3530. Bibcode:2011NJPh...13d3003S. doi:10.1088/1367-2630/13/4/043003. S2CID 118591567.
  17. ^ Łącki, M.; Zakrzewski, J. (2013). "Fast Dynamics for Atoms in Optical Lattices". Physical Review Letters. 110 (6): 065301. arXiv:1210.7957. Bibcode:2013PhRvL.110f5301L. doi:10.1103/PhysRevLett.110.065301. PMID 23432268. S2CID 29095052.
  18. ^ Will, S.; Best, T.; Schneider, U.; Hackermüller, L.; Lühmann, D. S. R.; Bloch, I. (2010). "Time-resolved observation of coherent multi-body interactions in quantum phase revivals". Nature. 465 (7295): 197–201. Bibcode:2010Natur.465..197W. doi:10.1038/nature09036. PMID 20463733. S2CID 4382706.
  19. ^ Bakr, Waseem S.; Gillen, Jonathon I.; Peng, Amy; Fölling, Simon; Greiner, Markus (2009). "A quantum gas microscope for detecting single atoms in a Hubbard-regime optical lattice". Nature. 462 (7269): 74–77. arXiv:0908.0174. Bibcode:2009Natur.462...74B. doi:10.1038/nature08482. PMID 19890326. S2CID 4419426.
  20. ^ Bakr, W. S.; Peng, A.; Tai, M. E.; Ma, R.; Simon, J.; Gillen, J. I.; Fölling, S.; Pollet, L.; Greiner, M. (2010-07-30). "Probing the Superfluid–to–Mott Insulator Transition at the Single-Atom Level". Science. 329 (5991): 547–550. arXiv:1006.0754. Bibcode:2010Sci...329..547B. doi:10.1126/science.1192368. ISSN 0036-8075. PMID 20558666. S2CID 3778258.
  21. ^ Weitenberg, Christof; Endres, Manuel; Sherson, Jacob F.; Cheneau, Marc; Schauß, Peter; Fukuhara, Takeshi; Bloch, Immanuel; Kuhr, Stefan (2011). "Single-spin addressing in an atomic Mott insulator". Nature. 471 (7338): 319–324. arXiv:1101.2076. Bibcode:2011Natur.471..319W. doi:10.1038/nature09827. PMID 21412333. S2CID 4352129.
  22. ^ Romero-Isart, O; Eckert, K; Rodó, C; Sanpera, A (2007). "Transport and entanglement generation in the Bose–Hubbard model". Journal of Physics A: Mathematical and Theoretical. 40 (28): 8019–31. arXiv:quant-ph/0703177. Bibcode:2007JPhA...40.8019R. doi:10.1088/1751-8113/40/28/S11. S2CID 11673450.
  23. ^ Jordan, J; Orus, R; Vidal, G (2009). "Numerical study of the hard-core Bose-Hubbard model on an infinite square lattice". Phys. Rev. B. 79 (17): 174515. arXiv:0901.0420. Bibcode:2009PhRvB..79q4515J. doi:10.1103/PhysRevB.79.174515. S2CID 119073171.
  24. ^ Kshetrimayum, A.; Rizzi, M.; Eisert, J.; Orus, R. (2019). "Tensor Network Annealing Algorithm for Two-Dimensional Thermal States". Phys. Rev. Lett. 122 (7): 070502. arXiv:1809.08258. Bibcode:2019PhRvL.122g0502K. doi:10.1103/PhysRevLett.122.070502. PMID 30848636. S2CID 53125536.
  25. ^ Eisert, J.; Cramer, M.; Plenio, M. B. (2010). "Colloquium: Area laws for the entanglement entropy". Reviews of Modern Physics. 82 (1): 277. arXiv:0808.3773. Bibcode:2010RvMP...82..277E. doi:10.1103/RevModPhys.82.277.
  26. ^ Góral, K.; Santos, L.; Lewenstein, M. (2002). "Quantum Phases of Dipolar Bosons in Optical Lattices". Physical Review Letters. 88 (17): 170406. arXiv:cond-mat/0112363. Bibcode:2002PhRvL..88q0406G. doi:10.1103/PhysRevLett.88.170406. PMID 12005738. S2CID 41827359.
  27. ^ Sowiński, T.; Dutta, O.; Hauke, P.; Tagliacozzo, L.; Lewenstein, M. (2012). "Dipolar Molecules in Optical Lattices". Physical Review Letters. 108 (11): 115301. arXiv:1109.4782. Bibcode:2012PhRvL.108k5301S. doi:10.1103/PhysRevLett.108.115301. PMID 22540482. S2CID 5438190.
  28. ^ Tsuchiya, S.; Kurihara, S.; Kimura, T. (2004). "Superfluid–Mott insulator transition of spin-1 bosons in an optical lattice". Physical Review A. 70 (4): 043628. arXiv:cond-mat/0209676. Bibcode:2004PhRvA..70d3628T. doi:10.1103/PhysRevA.70.043628. S2CID 118566913.
  29. ^ Gurarie, V.; Pollet, L.; Prokof’Ev, N. V.; Svistunov, B. V.; Troyer, M. (2009). "Phase diagram of the disordered Bose-Hubbard model". Physical Review B. 80 (21): 214519. arXiv:0909.4593. Bibcode:2009PhRvB..80u4519G. doi:10.1103/PhysRevB.80.214519. S2CID 54075502.

bose, hubbard, model, gives, description, physics, interacting, spinless, bosons, lattice, closely, related, hubbard, model, that, originated, solid, state, physics, approximate, description, superconducting, systems, motion, electrons, between, atoms, crystal. The Bose Hubbard model gives a description of the physics of interacting spinless bosons on a lattice It is closely related to the Hubbard model that originated in solid state physics as an approximate description of superconducting systems and the motion of electrons between the atoms of a crystalline solid The model was introduced by Gersch and Knollman 1 in 1963 in the context of granular superconductors The term Bose in its name refers to the fact that the particles in the system are bosonic The model rose to prominence in the 1980s after it was found to capture the essence of the superfluid insulator transition in a way that was much more mathematically tractable than fermionic metal insulator models 2 3 4 The Bose Hubbard model can be used to describe physical systems such as bosonic atoms in an optical lattice 5 as well as certain magnetic insulators 6 7 Furthermore it can be generalized and applied to Bose Fermi mixtures in which case the corresponding Hamiltonian is called the Bose Fermi Hubbard Hamiltonian Contents 1 Hamiltonian 1 1 Hilbert space 2 Phase diagram 3 Mean field theory 4 Implementation in optical lattices 4 1 Subtleties and approximations 5 Experimental results 6 Further applications 7 Numerical simulation 8 Generalizations 9 See also 10 ReferencesHamiltonian editThe physics of this model is given by the Bose Hubbard Hamiltonian H t i j b i b j b j b i U 2 i n i n i 1 m i n i displaystyle H t sum left langle i j right rangle left hat b i dagger hat b j hat b j dagger hat b i right frac U 2 sum i hat n i left hat n i 1 right mu sum i hat n i nbsp Here i j displaystyle left langle i j right rangle nbsp denotes summation over all neighboring lattice sites i displaystyle i nbsp and j displaystyle j nbsp while b i displaystyle hat b i dagger nbsp and b i displaystyle hat b i nbsp are bosonic creation and annihilation operators such that n i b i b i displaystyle hat n i hat b i dagger hat b i nbsp gives the number of particles on site i displaystyle i nbsp The model is parametrized by the hopping amplitude t displaystyle t nbsp that describes boson mobility in the lattice the on site interaction U displaystyle U nbsp which can be attractive U lt 0 displaystyle U lt 0 nbsp or repulsive U gt 0 displaystyle U gt 0 nbsp and the chemical potential m displaystyle mu nbsp which essentially sets the number of particles If unspecified typically the phrase Bose Hubbard model refers to the case where the on site interaction is repulsive This Hamiltonian has a global U 1 displaystyle U 1 nbsp symmetry which means that it is invariant its physical properties are unchanged by the transformation b i e i 8 b i displaystyle hat b i rightarrow e i theta hat b i nbsp In a superfluid phase this symmetry is spontaneously broken Hilbert space edit The dimension of the Hilbert space of the Bose Hubbard model is given by D b N b L 1 N b L 1 displaystyle D b N b L 1 N b L 1 nbsp where N b displaystyle N b nbsp is the total number of particles while L displaystyle L nbsp denotes the total number of lattice sites At fixed N b displaystyle N b nbsp or L displaystyle L nbsp the Hilbert space dimension D b displaystyle D b nbsp grows polynomially but at a fixed density of n b displaystyle n b nbsp bosons per site it grows exponentially as D b 1 n b 1 1 n b n b L textstyle D b sim left 1 n b left 1 frac 1 n b right n b right L nbsp Analogous Hamiltonians may be formulated to describe spinless fermions the Fermi Hubbard model or mixtures of different atom species Bose Fermi mixtures for example In the case of a mixture the Hilbert space is simply the tensor product of the Hilbert spaces of the individual species Typically additional terms are included to model interaction between species Phase diagram editAt zero temperature the Bose Hubbard model in the absence of disorder is in either a Mott insulating state at small t U displaystyle t U nbsp or in a superfluid state at large t U displaystyle t U nbsp 8 The Mott insulating phases are characterized by integer boson densities by the existence of an energy gap for particle hole excitations and by zero compressibility The superfluid is characterized by long range phase coherence a spontaneous breaking of the Hamiltonian s continuous U 1 displaystyle U 1 nbsp symmetry a non zero compressibility and superfluid susceptibility At non zero temperature in certain parameter regimes a regular fluid phase appears that does not break the U 1 displaystyle U 1 nbsp symmetry and does not display phase coherence Both of these phases have been experimentally observed in ultracold atomic gases 9 In the presence of disorder a third Bose glass phase exists 4 The Bose glass is a Griffiths phase and can be thought of as a Mott insulator containing rare puddles of superfluid These superfluid pools are not interconnected so the system remains insulating but their presence significantly changes model thermodynamics The Bose glass phase is characterized by finite compressibility the absence of a gap and by an infinite superfluid susceptibility 4 It is insulating despite the absence of a gap as low tunneling prevents the generation of excitations which although close in energy are spatially separated The Bose glass has a non zero Edwards Anderson order parameter 10 11 and has been suggested but not proven to display replica symmetry breaking 12 Mean field theory editThe phases of the clean Bose Hubbard model can be described using a mean field Hamiltonian 13 H MF i m n i U 2 n i n i 1 z t ps b i ps b i z t ps ps displaystyle begin aligned H textrm MF amp sum i left mu hat n i frac U 2 hat n i hat n i 1 zt psi hat b i psi hat b i dagger zt psi psi right end aligned nbsp where z displaystyle z nbsp is the lattice co ordination number This can be obtained from the full Bose Hubbard Hamiltonian by setting b i ps d b displaystyle hat b i rightarrow psi delta hat b nbsp where ps b i displaystyle psi langle hat b i rangle nbsp neglecting terms quadratic in d b i displaystyle delta hat b i nbsp assumedly infinitesimal and relabelling d b i b i displaystyle delta hat b i rightarrow hat b i nbsp Because this decoupling breaks the U 1 displaystyle U 1 nbsp symmetry of the initial Hamiltonian for all non zero values of ps displaystyle psi nbsp this parameter acts as a superfluid order parameter For simplicity this decoupling assumes ps displaystyle psi nbsp to be the same on every site which precludes exotic phases such as supersolids or other inhomogeneous phases Other decouplings are possible The phase diagram can be determined by calculating the energy of this mean field Hamiltonian using second order perturbation theory and finding the condition for which ps 0 displaystyle psi neq 0 nbsp To do this the Hamiltonian is written as a site local piece plus a perturbation H MF i h i 0 z t ps b i ps b i with h i 0 m n i U 2 n i n i 1 z t ps ps displaystyle H textrm MF sum i left h i 0 zt psi hat b i psi hat b i dagger right quad textrm with quad h i 0 mu hat n i frac U 2 hat n i hat n i 1 zt psi psi nbsp where the bilinear terms ps b i displaystyle psi hat b i nbsp and its conjugate are treated as the perturbation The order parameter ps displaystyle psi nbsp is assumed to be small near the phase transition The local term is diagonal in the Fock basis giving the zeroth order energy contribution E m 0 m m U 2 m m 1 z t ps 2 displaystyle E m 0 mu m frac U 2 m m 1 zt psi 2 nbsp where m displaystyle m nbsp is an integer that labels the filling of the Fock state The perturbative piece can be treated with second order perturbation theory which leads to E m 2 z t ps 2 n m m b i b i n 2 E n 0 E m 0 z t 2 ps 2 m U m 1 m m 1 m U m displaystyle E m 2 zt psi 2 sum n neq m frac langle m hat b i dagger hat b i n rangle 2 E n 0 E m 0 zt 2 psi 2 left frac m U m 1 mu frac m 1 mu Um right nbsp The energy can be expressed as a series expansion in even powers of the order parameter also known as the Landau formalism E constant R ps 2 W ps 4 displaystyle E text constant R psi 2 W psi 4 nbsp After doing so the condition for the mean field second order phase transition between the Mott insulator and the superfluid phase is given by r R z t 1 z t m U m 1 m m 1 m U m 0 displaystyle r frac R zt 1 zt left frac m U m 1 mu frac m 1 mu Um right 0 nbsp where the integer m displaystyle m nbsp describes the filling of the m t h displaystyle m th nbsp Mott insulating lobe Plotting the line r 0 displaystyle r 0 nbsp for different integer values of m displaystyle m nbsp generates the boundary of the different Mott lobes as shown in the phase diagram 4 Implementation in optical lattices editUltracold atoms in optical lattices are considered a standard realization of the Bose Hubbard model The ability to tune model parameters using simple experimental techniques and the lack of the lattice dynamics that are present in solid state electronic systems mean that ultracold atoms offer a clean controllable realisation of the Bose Hubbard model 14 5 The biggest downside with optical lattice technology is the trap lifetime with atoms typically trapped for only a few tens of seconds To see why ultracold atoms offer such a convenient realization of Bose Hubbard physics the Bose Hubbard Hamiltonian can be derived starting from the second quantized Hamiltonian that describes a gas of ultracold atoms in the optical lattice potential This Hamiltonian is given by H d 3 r ps r ℏ 2 2 m 2 V l a t t r ps r g 2 ps r ps r ps r ps r m ps r ps r displaystyle H int rm d 3 r left hat psi dagger vec r left frac hbar 2 2m nabla 2 V rm latt vec r right hat psi vec r frac g 2 hat psi dagger vec r hat psi dagger vec r hat psi vec r hat psi vec r mu hat psi dagger vec r hat psi vec r right nbsp where V l a t t displaystyle V latt nbsp is the optical lattice potential g displaystyle g nbsp is the contact interaction amplitude and m displaystyle mu nbsp is the chemical potential The tight binding approximation results in the substitution ps r i w i a r b i a textstyle hat psi vec r sum limits i w i alpha vec r b i alpha nbsp which leads to the Bose Hubbard Hamiltonian the physics are restricted to the lowest band a 0 displaystyle alpha 0 nbsp and the interactions are local at the level of the discrete mode Mathematically this can be stated as the requirement that w i a r w j b r w k g r w l d r d 3 r 0 textstyle int w i alpha vec r w j beta vec r w k gamma r w l delta vec r rm d 3 r 0 nbsp except for case i j k l a b g d 0 displaystyle i j k l wedge alpha beta gamma delta 0 nbsp Here w i a r displaystyle w i alpha vec r nbsp is a Wannier function for a particle in an optical lattice potential localized around site i displaystyle i nbsp of the lattice and for the a displaystyle alpha nbsp th Bloch band 15 Subtleties and approximations edit The tight binding approximation significantly simplifies the second quantized Hamiltonian though it introduces several limitations at the same time For single site states with several particles in a single state the interactions may couple to higher Bloch bands which contradicts base assumptions Still a single band model is able to address low energy physics of such a setting but with parameters U and J becoming density dependent Instead of one parameter U the interaction energy of n particles may be described by U n displaystyle U n nbsp close but not equal to U 15 When considering fast lattice dynamics additional terms are added to the Hamiltonian so that the time dependent Schrodinger equation is obeyed in the time dependent Wannier function basis The terms come from the Wannier functions time dependence 16 17 Otherwise the lattice dynamics may be incorporated by making the key parameters of the model time dependent varying with the instantaneous value of the optical potential Experimental results editQuantum phase transitions in the Bose Hubbard model were experimentally observed by Greiner et al 9 and density dependent interaction parameters U n displaystyle U n nbsp were observed by Immanuel Bloch s group 18 Single atom resolution imaging of the Bose Hubbard model has been possible since 2009 using quantum gas microscopes 19 20 21 Further applications editThe Bose Hubbard model is of interest in the field of quantum computation and quantum information Entanglement of ultra cold atoms can be studied using this model 22 Numerical simulation editIn the calculation of low energy states the term proportional to n 2 U displaystyle n 2 U nbsp means that large occupation of a single site is improbable allowing for truncation of local Hilbert space to states containing at most d lt displaystyle d lt infty nbsp particles Then the local Hilbert space dimension is d 1 displaystyle d 1 nbsp The dimension of the full Hilbert space grows exponentially with the number of lattice sites limiting exact computer simulations of the entire Hilbert space to systems of 15 20 particles in 15 20 lattice sites citation needed Experimental systems contain several million sites with average filling above unity citation needed One dimensional lattices may be studied using density matrix renormalization group DMRG and related techniques such as time evolving block decimation TEBD This includes calculating the ground state of the Hamiltonian for systems of thousands of particles on thousands of lattice sites and simulating its dynamics governed by the time dependent Schrodinger equation Recently when two dimensional lattices have been studied using projected entangled pair states a generalization of matrix product states in higher dimensions both for the ground state 23 and finite temperature 24 Higher dimensions are significantly more difficult due to the rapid growth of entanglement 25 All dimensions may be treated by quantum Monte Carlo algorithms citation needed which provide a way to study properties of the Hamiltonian s thermal states and in particular the ground state Generalizations editBose Hubbard like Hamiltonians may be derived for different physical systems containing ultracold atom gas in the periodic potential They include systems with longer ranged density density interactions of the form V n i n j displaystyle Vn i n j nbsp which may stabilise a supersolid phase for certain parameter values dimerised magnets where spin 1 2 electrons are bound together in pairs called dimers that have bosonic excitation statistics and are described by a Bose Hubbard model long range dipolar interaction 26 systems with interaction induced tunneling terms a i n i n j a j displaystyle a i dagger n i n j a j nbsp 27 internal spin structure of atoms for example due to trapping an entire degenerate manifold of hyperfine spin states for F 1 it leads to the spin 1 Bose Hubbard model 28 clarification needed situations where the gas experiences an additional potential for example in disordered systems 29 The disorder might be realised by a speckle pattern or using a second incommensurate weaker optical lattice In the latter case inclusion of the disorder amounts to including extra term of the form H I V d i cos k i f n i displaystyle H I V d sum limits i cos ki varphi hat n i nbsp See also editJaynes Cummings Hubbard modelReferences edit Gersch H Knollman G 1963 Quantum Cell Model for Bosons Physical Review 129 2 959 Bibcode 1963PhRv 129 959G doi 10 1103 PhysRev 129 959 Ma M Halperin B I Lee P A 1986 09 01 Strongly disordered superfluids Quantum fluctuations and critical behavior Physical Review B 34 5 3136 3143 Bibcode 1986PhRvB 34 3136M doi 10 1103 PhysRevB 34 3136 PMID 9940047 Giamarchi T Schulz H J 1988 01 01 Anderson localization and interactions in one dimensional metals Physical Review B 37 1 325 340 Bibcode 1988PhRvB 37 325G doi 10 1103 PhysRevB 37 325 PMID 9943580 a b c d Fisher Matthew P A Grinstein G Fisher Daniel S 1989 Boson localization and the superfluid insulator transition PDF Physical Review B 40 1 546 70 Bibcode 1989PhRvB 40 546F doi 10 1103 PhysRevB 40 546 PMID 9990946 a b Jaksch D Zoller P 2005 The cold atom Hubbard toolbox Annals of Physics 315 1 52 arXiv cond mat 0410614 Bibcode 2005AnPhy 315 52J CiteSeerX 10 1 1 305 9031 doi 10 1016 j aop 2004 09 010 S2CID 12352119 Giamarchi Thierry Ruegg Christian Tchernyshyov Oleg 2008 Bose Einstein condensation in magnetic insulators Nature Physics 4 3 198 204 arXiv 0712 2250 Bibcode 2008NatPh 4 198G doi 10 1038 nphys893 S2CID 118661914 Zapf Vivien Jaime Marcelo Batista C D 2014 05 15 Bose Einstein condensation in quantum magnets Reviews of Modern Physics 86 2 563 614 Bibcode 2014RvMP 86 563Z doi 10 1103 RevModPhys 86 563 Kuhner T Monien H 1998 Phases of the one dimensional Bose Hubbard model Physical Review B 58 22 R14741 arXiv cond mat 9712307 Bibcode 1998PhRvB 5814741K doi 10 1103 PhysRevB 58 R14741 S2CID 118911555 a b Greiner Markus Mandel Olaf Esslinger Tilman Hansch Theodor W Bloch Immanuel 2002 Quantum phase transition from a superfluid to a Mott insulator in a gas of ultracold atoms Nature 415 6867 39 44 Bibcode 2002Natur 415 39G doi 10 1038 415039a PMID 11780110 S2CID 4411344 Morrison S Kantian A Daley A J Katzgraber H G Lewenstein M Buchler H P Zoller P 2008 Physical replicas and the Bose glass in cold atomic gases New Journal of Physics 10 7 073032 arXiv 0805 0488 Bibcode 2008NJPh 10g3032M doi 10 1088 1367 2630 10 7 073032 ISSN 1367 2630 S2CID 2822103 Thomson S J Walker L S Harte T L Bruce G D 2016 11 03 Measuring the Edwards Anderson order parameter of the Bose glass A quantum gas microscope approach Physical Review A 94 5 051601 arXiv 1607 05254 Bibcode 2016PhRvA 94e1601T doi 10 1103 PhysRevA 94 051601 S2CID 56027819 Thomson S J Kruger F 2014 Replica symmetry breaking in the Bose glass EPL 108 3 30002 arXiv 1312 0515 Bibcode 2014EL 10830002T doi 10 1209 0295 5075 108 30002 S2CID 56307253 Sachdev Subir 2011 Quantum phase transitions Cambridge University Press ISBN 9780521514682 OCLC 693207153 Jaksch D Bruder C Cirac J Gardiner C Zoller P 1998 Cold Bosonic Atoms in Optical Lattices Physical Review Letters 81 15 3108 arXiv cond mat 9805329 Bibcode 1998PhRvL 81 3108J doi 10 1103 PhysRevLett 81 3108 S2CID 55578669 a b Luhmann D S R Jurgensen O Sengstock K 2012 Multi orbital and density induced tunneling of bosons in optical lattices New Journal of Physics 14 3 033021 arXiv 1108 3013 Bibcode 2012NJPh 14c3021L doi 10 1088 1367 2630 14 3 033021 S2CID 119216031 Sakmann K Streltsov A I Alon O E Cederbaum L S 2011 Optimal time dependent lattice models for nonequilibrium dynamics New Journal of Physics 13 4 043003 arXiv 1006 3530 Bibcode 2011NJPh 13d3003S doi 10 1088 1367 2630 13 4 043003 S2CID 118591567 Lacki M Zakrzewski J 2013 Fast Dynamics for Atoms in Optical Lattices Physical Review Letters 110 6 065301 arXiv 1210 7957 Bibcode 2013PhRvL 110f5301L doi 10 1103 PhysRevLett 110 065301 PMID 23432268 S2CID 29095052 Will S Best T Schneider U Hackermuller L Luhmann D S R Bloch I 2010 Time resolved observation of coherent multi body interactions in quantum phase revivals Nature 465 7295 197 201 Bibcode 2010Natur 465 197W doi 10 1038 nature09036 PMID 20463733 S2CID 4382706 Bakr Waseem S Gillen Jonathon I Peng Amy Folling Simon Greiner Markus 2009 A quantum gas microscope for detecting single atoms in a Hubbard regime optical lattice Nature 462 7269 74 77 arXiv 0908 0174 Bibcode 2009Natur 462 74B doi 10 1038 nature08482 PMID 19890326 S2CID 4419426 Bakr W S Peng A Tai M E Ma R Simon J Gillen J I Folling S Pollet L Greiner M 2010 07 30 Probing the Superfluid to Mott Insulator Transition at the Single Atom Level Science 329 5991 547 550 arXiv 1006 0754 Bibcode 2010Sci 329 547B doi 10 1126 science 1192368 ISSN 0036 8075 PMID 20558666 S2CID 3778258 Weitenberg Christof Endres Manuel Sherson Jacob F Cheneau Marc Schauss Peter Fukuhara Takeshi Bloch Immanuel Kuhr Stefan 2011 Single spin addressing in an atomic Mott insulator Nature 471 7338 319 324 arXiv 1101 2076 Bibcode 2011Natur 471 319W doi 10 1038 nature09827 PMID 21412333 S2CID 4352129 Romero Isart O Eckert K Rodo C Sanpera A 2007 Transport and entanglement generation in the Bose Hubbard model Journal of Physics A Mathematical and Theoretical 40 28 8019 31 arXiv quant ph 0703177 Bibcode 2007JPhA 40 8019R doi 10 1088 1751 8113 40 28 S11 S2CID 11673450 Jordan J Orus R Vidal G 2009 Numerical study of the hard core Bose Hubbard model on an infinite square lattice Phys Rev B 79 17 174515 arXiv 0901 0420 Bibcode 2009PhRvB 79q4515J doi 10 1103 PhysRevB 79 174515 S2CID 119073171 Kshetrimayum A Rizzi M Eisert J Orus R 2019 Tensor Network Annealing Algorithm for Two Dimensional Thermal States Phys Rev Lett 122 7 070502 arXiv 1809 08258 Bibcode 2019PhRvL 122g0502K doi 10 1103 PhysRevLett 122 070502 PMID 30848636 S2CID 53125536 Eisert J Cramer M Plenio M B 2010 Colloquium Area laws for the entanglement entropy Reviews of Modern Physics 82 1 277 arXiv 0808 3773 Bibcode 2010RvMP 82 277E doi 10 1103 RevModPhys 82 277 Goral K Santos L Lewenstein M 2002 Quantum Phases of Dipolar Bosons in Optical Lattices Physical Review Letters 88 17 170406 arXiv cond mat 0112363 Bibcode 2002PhRvL 88q0406G doi 10 1103 PhysRevLett 88 170406 PMID 12005738 S2CID 41827359 Sowinski T Dutta O Hauke P Tagliacozzo L Lewenstein M 2012 Dipolar Molecules in Optical Lattices Physical Review Letters 108 11 115301 arXiv 1109 4782 Bibcode 2012PhRvL 108k5301S doi 10 1103 PhysRevLett 108 115301 PMID 22540482 S2CID 5438190 Tsuchiya S Kurihara S Kimura T 2004 Superfluid Mott insulator transition of spin 1 bosons in an optical lattice Physical Review A 70 4 043628 arXiv cond mat 0209676 Bibcode 2004PhRvA 70d3628T doi 10 1103 PhysRevA 70 043628 S2CID 118566913 Gurarie V Pollet L Prokof Ev N V Svistunov B V Troyer M 2009 Phase diagram of the disordered Bose Hubbard model Physical Review B 80 21 214519 arXiv 0909 4593 Bibcode 2009PhRvB 80u4519G doi 10 1103 PhysRevB 80 214519 S2CID 54075502 Retrieved from https en wikipedia org w index php title Bose Hubbard model amp oldid 1151371311, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.