Viewpoint

Artificial magnetism for ultracold atoms

    Gediminas Juzeliūnas
    • Institute of Theoretical Physics and Astronomy of Vilnius University, A. Goštauto 12, Vilnius 01108, Lithuania
Physics 2, 25
Trapped cold atom gases mimic much of the behavior of electrons in a solid, but because the atoms are neutral, it is difficult to imitate the physics of electrons moving in a magnetic field. Now, experiments show that a suitable combination of lasers can create an artificial magnetic field for cold atoms.
Figure 1: Energy levels of an atom with a total spin F=1 containing three ground states with different spin projections mF=0,±1. Two counter-propagating laser beams induce Raman transitions between these states via the absorption (emission) of a photon with a wave vector kr from one beam, and emission (absorption) of a photon from another beam with the opposite wave vector krkr. The laser beams couple the atomic states |mF,k differing in the spin projection ΔmF=1 and the linear momentum Δk=2kr , the latter 2kr describing the two-photon recoil kick . This leads to the formation of the dressed states representing a superposition of the atomic states with the different spin projections and different center-of-mass momenta. If the dressed states have a spatial dependence in the direction orthogonal to kr, the atomic motion in that direction will be accompanied by transverse recoil kicks due to the population redistribution in the spin components comprising the dressed state. This will simulate the Lorentz force. A more detail discussion on the semiclassical interpretation of the light-induced artificial Lorentz force is presented in Ref. [20].

Recent progress in cooling and trapping of atoms has stimulated intense studies of the physical properties of the atomic Bose-Einstein condensates (BECs) and the degenerate Fermi gases. These systems are formed typically at densities of 1014 atoms per cm3 and temperatures in the nano-Kelvin range. Much current research in atomic quantum gases is motivated by the possibility of simulating a variety of condensed-matter phenomena [1–4] such as the transition between a superfluid state and an insulating state. An advantage of using atomic systems is the high degree of control: One can relatively easily change the physical parameters of the system including the number of the trapped atoms, the shape of the trapping potential, and the strength of the atom-atom coupling.

On the other hand, the atoms forming the quantum gases are electrically neutral particles and there is no Lorentz force acting on them. Thus there is no direct analogy between the properties of the degenerate atomic gases and magnetic phenomena involving electrons in solids, such as the quantum Hall effect discussed in a recent Viewpoint [5]. Now, Yu-Ju Lin and colleagues at NIST in Gaithersburg, Maryland, report in Physical Review Letters an important experimental advance towards producing an artificial magnetic field for an atomic Bose-Einstein condensate [6].

The usual way to imitate the magnetic field in a cloud of electrically neutral atoms is to rotate the system [7–9]. In the rotating frame of reference the Hamiltonian for the atomic motion acquires a vector-potential-type term describing the Coriolis force. The latter has the same mathematical structure as the Lorentz force for a charged particle in a uniform magnetic field. Additionally, inertial effects push the atoms away from the center of the rotating cloud. When the rotation frequency approaches the frequency of the atomic trap, the trapping is canceled by the centrifugal effect and the problem reduces to the cyclotron motion of a charged particle in a constant magnetic field. Using this method it is possible to produce vortex lattices in the atomic Bose-Einstein condensates [1].

Limitations in the experimental techniques have so far prevented researchers from reaching higher artificial magnetic fields where cold atoms could enter the fractional quantum Hall regime. Furthermore, the method of rotation usually applies to atomic clouds confined by cylindrically symmetric traps with a small anisotropic rotating potential added [1,7–9]. However, there are important trapping configurations like atom chips [10] lacking axial symmetry. Therefore it is desirable to have alternative methods for producing artificial magnetic fields for cold atoms without rotation. For atoms trapped in an optical lattice, the artificial magnetic field can be generated by inducing an asymmetry in the atomic tunneling between the lattice sites [11–13]. To simulate a magnetic flux there should be a nonvanishing phase for the atomic transfer around an elementary cell of the lattice known as the Peierls phase.

It is possible to create an artificial magnetic field for cold atoms without a trapping lattice. In these proposals, several laser beams are used to induce transitions between different atomic internal states in a position-dependent manner [14–17]. The laser fields alter (“dress”) the internal atomic eigenstates, which become superpositions of the original atomic states. It is the position-dependence of these dressed eigenstates that changes the momentum operator p=-i ħ to p-A for the atomic center-of-mass motion in the nth internal state [18,19]. The appearing effective vector potential A represents an atomic momentum associated with the nth internal dressed state. Adopting the adiabatic approximation, one can ignore transitions to other atomic levels. The atomic motion in the nth dressed state then becomes equivalent to a motion of a particle of a unit charge in the magnetic field B=×A. Several schemes have been proposed providing a nonzero effective magnetic field B for electronically neutral atoms in the laser fields [15–17]. Yet up to now there has been no experimental evidence of generating the artificial magnetic field using methods other than rotation.

The experiment by Y.-J. Lin et al. [6] makes use of two counter-propagating laser beams that act on a BEC of 87Rb atoms characterized by a total spin F=1. The atoms contain three magnetic ground states |mF corresponding to different spin projections mF=0,±1. Two laser beams induce the Raman transitions between the magnetic states with ΔmF=1. The Raman process involves the absorption (emission) of a photon with a wave vector kr from one beam, and emission (absorption) of a photon from another beam with the opposite wave vector kr-kr as shown in Fig. 1. As a result the laser beams couple the atomic states |mF,k differing in spin projection ΔmF=1 and the linear momentum Δk=2kr . In such a situation the laser-dressed atomic states are the superpositions of the combined spin and the center-of-mass motion states |k=|-1,k-2krc-1+|0,kc0+|1,k+2krc1, where cj is the probability amplitude for the atom to have the spin projection mF=j. The states |k have a minimum energy at a certain wave vector k=kmin, which depends on the detuning from the two-photon resonance. The corresponding momentum ħ kmin plays a role of a spatially uniform effective vector potential, shifting the origin of the atomic momentum. To simulate a Lorentz force F=v×B, the effective vector potential A= ħ kmin should be spatially dependent and have a nonzero curl: B=×A0.

An important ingredient of the experiment by Y.-J. Lin et al. [6] is the application of a real magnetic field in addition to the laser beams. The magnetic field removes the initial degeneracy of the spin levels with mF=0,±1 via the linear and quadratic Zeeman effects. By changing the strength of the magnetic field one can increase or decrease the two-photon detuning δ and thus alter the effective vector potential A= ħ kmin(δ). Lin et al. have managed to transfer adiabatically the condensate atoms to the state with k=kmin(δ) in the presence of the Raman laser beams and the magnetic field. Subsequently, the condensate was allowed to expand by removing the trap. The time-of-flight measurements have determined the atomic momentum ħ kmin(δ) as a function of the detuning δ, showing a good agreement with the theoretical calculations [6].

The vector potential A= ħ kmin(δ) produced in this way is uniform and thus corresponds to a zero effective magnetic field: B=×A=0 [6]. The technique can be extended to generate nonuniform vector potentials for cold atoms. To accomplish this Lin et al. propose to apply a nonhomogeneous real magnetic field [6], resulting in the distance-dependent detuning δ. This provides the spatial dependence of the wave vector kmin(δ) and leads to the nonzero effective magnetic field B= ħ kmin(δ)×kmin. In this way, it is a combination of the inhomogeneous real magnetic field and the counter-propagating laser beams that results in the artificial magnetic field and the Lorentz force acting on the electrically neutral atoms. The suggested scheme [6] can serve as an alternative to the previous proposals relying on the spatial variation of the laser beam profiles [15–17] or the optical lattices [11–13]. The future will show which of these approaches works better in creating the artificial magnetism.

If this could be achieved, we would have within our grasp the experimental realization of gauge potentials for cold atoms that would allow us to properly simulate a range of fascinating condensed-matter and high-energy physics. An important issue to be explored would be the creation of the non-Abelian gauge potentials for cold atoms [21–26]. The non-Abelian effects can appear if the atoms have degenerate dressed states and hence the atomic center-of-mass motion is described by a multicomponent wave function. The light-induced vector potential becomes then a matrix, the Cartesian components of which do not necessarily commute. By choosing the proper light fields, one can simulate a number of intriguing phenomena, such as formation of non-Abelian magnetic monopoles for cold atoms [22,26].

References

  1. I. Bloch, J. Dalibard, and W. Zwerger, Rev. Mod. Phys. 80, 885 (2008)
  2. M. Greiner, O. Mandel, T. Esslinger, T. Hänsch, and I. Bloch, Nature 415, 39 (2002)
  3. Z. Hadzibabic, P. Krüger, M. Cheneau, B. Battelier, and J. Dalibard, Nature 441, 1118 (2006)
  4. M. Greiner, C. A. Regal, and D. S. Jin, Phys. Rev. Lett. 94, 110401 (2005)
  5. H. A. Fertig, Physics 2, 15 (2009)
  6. Y-J. Lin, R. L. Compton, A. R. Perry, W. D. Phillips, J. V. Porto, and I. B. Spielman, Phys. Rev. Lett. 102, 130401 (2009)
  7. K. W. Madison, F. Chevy, W. Wohlleben, and J. Dalibard, Phys. Rev. Lett. 84, 806 (2000)
  8. J. R. Abo-Shaeer, C. Raman, J. M. Vogels, and W. Ketterle, Science 292, 476 (2001)
  9. E. Hodby, G. Hechenblaikner, S. A. Hopkins, O. M. Marago, and C. J. Foot, Phys. Rev. Lett. 88, 010405 (2002)
  10. R. Folman, P. Krüger, J. Schmiedmayer, J. Denschlag, and C. Henkel, Adv. At. Mol. Opt. Phys. 48, 263 (2002)
  11. D. Jaksch and P. Zoller, New J. Phys. 5, 561 (2003)
  12. E. J. Mueller, Phys. Rev. A 70, 041603 (2004)
  13. A. S. Sørensen, E. Demler, and M. D. Lukin, Phys. Rev. Lett. 94, 086803 (2005)
  14. R. Dum and M. Olshanii, Phys. Rev. Lett. 76, 1788 (1996)
  15. G. Juzeliūnas and P. Öhberg, Phys. Rev. Lett. 93, 033602 (2004)
  16. G. Juzeliūnas, J. Ruseckas, P. Öhberg, and M.Fleischhauer, Phys. Rev. A 73, 025602 (2006)
  17. K. J. Günter, M. Cheneau, T. Yefsah, S. P. Rath, and J. Dalibard, Phys. Rev. A 79, 011604 (2009)
  18. F. Wilczek and A. Shapere, Geometric Phases in Physics (World Scientific, Singapore, 1989)[Amazon][WorldCat]
  19. A. Bohm, A. Mostafazadeh, H. Koizumi, Q. Niu, and J. Zwanziger, The Geometric Phase in Quantum Systems (Springer, New York, 2003)[Amazon][WorldCat]
  20. M. Cheneau, S. P. Rath, T. Yefsah, K. J. Günter, G. Juzeliūnas, and J. Dalibard, Europhys. Lett. 83, 60001 (2008)
  21. K. Osterloh, M. Baig, L. Santos, P. Zoller, and M. Lewenstein, Phys. Rev. Lett. 95, 010403 (2005)
  22. J. Ruseckas, G. Juzeliūnas, P. Öhberg, and M. Fleischhauer, Phys. Rev. Lett. 95, 010404 (2005)
  23. J. Y. Vaishnav and C. W. Clark, Phys. Rev. Lett. 100, 153002 (2008)
  24. G. Juzeliūnas, J. Ruseckas, A. Jacob, L. Santos, and P. Öhberg, Phys. Rev. Lett. 100, 200405 (2008)
  25. N. Goldman, A. Kubasiak, P. Gaspard, and M. Lewenstein, Phys. Rev. A 79, 023624 (2009)
  26. V. Pietilä and M. Möttönen, Phys. Rev. Lett. 102, 080403 (2009)

About the Author

Image of Gediminas Juzeliūnas

Gediminas Juzeliūnas is a principal researcher and a deputy director at the Institute of Theoretical Physics and Astronomy of Vilnius University, Lithuania. He is also professor of physics at Lithuanian University of Educational Sciences in Vilnius. Dr. Juzeliūnas completed his Ph.D. in 1986 in theoretical condensed matter physics at Vilnius University, studying optical properties of excitons in confined geometries. Subsequently he held a two-year postdoctoral appointment at the University of East Anglia, England, shifting his research area towards quantum optics. Dr. Juzeliūnas was a Humboldt Research Fellow at the University of Ulm, Germany (1997–1998), and a Fulbright Scholar at the University of Oregon in the US (2000–2001). Dr. Juzeliūnas received a National State Prize for Science of Lithuania in 2008 and a Vilnius University Rector’s award in 2010. His current research focuses on ultracold atomic gases, slow light, metamaterials, and graphene. This includes a pioneering theoretical work on light-induced gauge potential for ultracold atoms.


Read PDF

Subject Areas

Atomic and Molecular Physics

Related Articles

Seeing Collisions in Cold Molecular Clouds
Atomic and Molecular Physics

Seeing Collisions in Cold Molecular Clouds

Dense ensembles of laser-cooled molecules allow the observation of molecular collisions—a result that could lead to applications of cold molecular gases in quantum simulation and fundamental physics tests. Read More »

Probing Liquid Water’s Structure with Attosecond X-Ray Pulses
Condensed Matter Physics

Probing Liquid Water’s Structure with Attosecond X-Ray Pulses

Using an ultrafast technique, researchers shed light on how the hydrogen-bonded structure of water is reflected in its x-ray spectrum. Read More »

Precise Measurement of Hydrogen’s Energy Levels
Atomic and Molecular Physics

Precise Measurement of Hydrogen’s Energy Levels

Researchers have measured the transition energy of several highly excited states, which could help resolve a discrepancy about the size of the proton. Read More »

More Articles