Optical lattice

Last updated

Atoms (represented as blue spheres) pictured in a 2D-optical lattice potential (represented as the yellow surface). AtomsInLattice.png
Atoms (represented as blue spheres) pictured in a 2D-optical lattice potential (represented as the yellow surface).

An optical lattice is formed by the interference of counter-propagating laser beams, creating a spatially periodic polarization pattern. The resulting periodic potential may trap neutral atoms via the Stark shift. [1] Atoms are cooled and congregate at the potential extrema (at maxima for blue-detuned lattices, and minima for red-detuned lattices). The resulting arrangement of trapped atoms resembles a crystal lattice [2] and can be used for quantum simulation.

Contents

Atoms trapped in the optical lattice may move due to quantum tunneling, even if the potential well depth of the lattice points exceeds the kinetic energy of the atoms, which is similar to the electrons in a conductor. [3] However, a superfluidMott insulator transition [4] may occur, if the interaction energy between the atoms becomes larger than the hopping energy when the well depth is very large. In the Mott insulator phase, atoms will be trapped in the potential minima and cannot move freely, which is similar to the electrons in an insulator. In the case of fermionic atoms, if the well depth is further increased the atoms are predicted to form an antiferromagnetic, i.e. Néel state at sufficiently low temperatures. [5]

History

Trapping atoms in standing waves of light was first proposed by V.S. Letokhov in 1968. [6]

Parameters

There are two important parameters of an optical lattice: the potential well depth and the periodicity.

Control of potential depth

The potential experienced by the atoms is related to the intensity of the laser used to generate the optical lattice. The potential depth of the optical lattice can be tuned in real time by changing the power of the laser, which is normally controlled by an acousto-optic modulator (AOM). The AOM is tuned to deflect a variable amount of the laser power into the optical lattice. Active power stabilization of the lattice laser can be accomplished by feedback of a photodiode signal to the AOM.

Control of periodicity

The periodicity of the optical lattice can be tuned by changing the wavelength of the laser or by changing the relative angle between the two laser beams. The real-time control of the periodicity of the lattice is still a challenging task. The wavelength of the laser cannot easily be varied over a large range in real time, and so the periodicity of the lattice is normally controlled by the relative angle between the laser beams. [7] However, it is difficult to keep the lattice stable while changing the relative angles, since the interference is sensitive to the relative phase between the laser beams. Titanium-sapphire lasers, with their large tunable range, provide a possible platform for direct tuning of wavelength in optical lattice systems.

Continuous control of the periodicity of a one-dimensional optical lattice while maintaining trapped atoms in-situ was first demonstrated in 2005 using a single-axis servo-controlled galvanometer. [8] This "accordion lattice" was able to vary the lattice periodicity from 1.30 to 9.3 μm. More recently, a different method of real-time control of the lattice periodicity was demonstrated, [9] in which the center fringe moved less than 2.7 μm while the lattice periodicity was changed from 0.96 to 11.2 μm. Keeping atoms (or other particles) trapped while changing the lattice periodicity remains to be tested more thoroughly experimentally. Such accordion lattices are useful for controlling ultracold atoms in optical lattices, where small spacing is essential for quantum tunneling, and large spacing enables single-site manipulation and spatially resolved detection. Site-resolved detection of the occupancy of lattice sites of both bosons and fermions within a high tunneling regime is regularly performed in quantum gas microscopes. [10] [11]

Principle of operation

A basic 1D optical lattice is formed by the interference pattern of two counter-propagating laser beams of the same linear polarization, most often in the far-detuned regime. The trapping mechanism is via the Stark shift, where off-resonant light causes shifts to an atom's internal structure. The effect of the Stark shift is to create a potential proportional to the intensity. This is the same trapping mechanism as in optical dipole traps (ODTs), with the only major difference being that the intensity of an optical lattice has a much more dramatic spatial variation than a standard ODT. [1]


The energy shift to (and thus, the potential experienced by) an electronic ground state is given by second-order time-independent perturbation theory, where the rapid time variation of the lattice potential at optical frequencies has been time-averaged.

where are the transition matrix elements for transitions from the ground state to the excited states . For a two-level system, this simplifies to

where is the linewidth of the excited state transition. [1]

An alternative picture of the stimulated light forces due to the AC Stark effect is to view the process as a stimulated Raman process, where the atom redistributes photons between the counterpropagating laser beams which form the lattice. In this picture, it is clearer that the atoms can only acquire momentum from the lattice in units of , where is the momentum of a photon of one laser beam. [1]

By use of additional laser beams, two- or three-dimensional optical lattices may be constructed. A 2D optical lattice may be constructed by interfering two orthogonal optical standing waves, giving rise to an array of 1D potential tubes. Likewise, three orthogonal optical standing waves can give rise to a 3D array of sites which may be approximated as tightly confining harmonic oscillator potentials. [2]

Technical challenges

The trapping potential experienced by atoms in an optical dipole trap is weak, generally below 1 mK. Thus atoms must be cooled significantly before loading them into the optical lattice. Cooling techniques used to this end include magneto-optical traps, Doppler cooling, polarization gradient cooling, Raman cooling, resolved sideband cooling, and evaporative cooling. [1]

Once cold atoms are loaded into the optical lattice, they will experience heating by various mechanisms such as spontaneous scattering of photons from the optical lattice lasers. These mechanisms generally limit the lifetime of optical lattice experiments. [1]

Time of flight imaging

Once cooled and trapped in an optical lattice, they can be manipulated or left to evolve. Common manipulations involve the "shaking" of the optical lattice by varying the relative phase between the counterpropagating beams or by modulating the frequency of one of the counterpropagating beams, or amplitude modulation of the lattice. After evolving in response to the lattice potential and any manipulations, the atoms can be imaged via absorption imaging.

A common observation technique is time of flight (TOF) imaging. TOF imaging works by first waiting some amount of time for the atoms to evolve in the lattice potential, then turning off the lattice potential (by switching off the laser power with an AOM). The atoms, now free, spread out at different rates according to their momenta. By controlling the amount of time the atoms are allowed to evolve, the distance travelled by atoms maps onto what their momentum state must have been when the lattice was turned off. Because the atoms in the lattice can only change in momentum by , a characteristic pattern in a TOF image of an optical-lattice system is a series of peaks along the lattice axis at momenta , where . Using TOF imaging, the momentum distribution of atoms in the lattice can be determined. Combined with in-situ absorption images (taken with the lattice still on), this is enough to determine the phase space density of the trapped atoms, an important metric for diagnosing Bose–Einstein condensation (or more generally, the formation of quantum degenerate phases of matter).

Uses

Quantum simulation

Atoms in an optical lattice provide an ideal quantum system where all parameters are highly controllable and where simplified models of condensed-matter physics may be experimentally realized. Because atoms can be imaged directly – something difficult to do with electrons in solids – they can be used to study effects that are difficult to observe in real crystals. Quantum gas microscopy techniques applied to trapped atom optical-lattice systems can even provide single-site imaging resolution of their evolution. [10]

By interfering differing numbers of beams in various geometries, varying lattice geometries can be created. These range from the simplest case of two counterpropagating beams forming a one-dimensional lattice, to more complex geometries like hexagonal lattices. The variety of geometries that can be produced in optical lattice systems allow the physical realization of different Hamiltonians, such as the Bose–Hubbard model, [4] the Kagome lattice and Sachdev–Ye–Kitaev model, [12] and the Aubry–André model. By studying the evolution of atoms under the influence of these Hamiltonians, insight about the solutions to the Hamiltonian can be gained. This is particularly relevant to complicated Hamiltonians which are not easily solvable using theoretical or numerical techniques, such as those for strongly correlated systems.

Optical clocks

The best atomic clocks in the world use atoms trapped in optical lattices, to obtain narrow spectral lines that are unaffected by the Doppler effect and recoil. [13] [14]

Quantum information

They are also promising candidates for quantum information processing. [15] [16]

Atom interferometry

Shaken optical lattices – where the phase of the lattice is modulated, causing the lattice pattern to scan back and forth – can be used to control the momentum state of the atoms trapped in the lattice. This control is exercised to split the atoms into populations of different momenta, propagate them to accumulate phase differences between the populations, and recombine them to produce an interference pattern. [17]

Other uses

Besides trapping cold atoms, optical lattices have been widely used in creating gratings and photonic crystals. They are also useful for sorting microscopic particles, [18] and may be useful for assembling cell arrays.

See also

Related Research Articles

<span class="mw-page-title-main">Bose–Einstein condensate</span> State of matter

In condensed matter physics, a Bose–Einstein condensate (BEC) is a state of matter that is typically formed when a gas of bosons at very low densities is cooled to temperatures very close to absolute zero. Under such conditions, a large fraction of bosons occupy the lowest quantum state, at which microscopic quantum-mechanical phenomena, particularly wavefunction interference, become apparent macroscopically. More generally, condensation refers to the appearance of macroscopic occupation of one or several states: for example, in BCS theory, a superconductor is a condensate of Cooper pairs. As such, condensation can be associated with phase transition, and the macroscopic occupation of the state is the order parameter.

<span class="mw-page-title-main">Laser cooling</span> Class of methods for cooling atoms to very low temperatures

Laser cooling includes several techniques where atoms, molecules, and small mechanical systems are cooled with laser light. The directed energy of lasers is often associated with heating materials, e.g. laser cutting, so it can be counterintuitive that laser cooling often results in sample temperatures approaching absolute zero. It is a routine step in many atomic physics experiments where the laser-cooled atoms are then subsequently manipulated and measured, or in technologies, such as atom-based quantum computing architectures. Laser cooling relies on the change in momentum when an object, such as an atom, absorbs and re-emits a photon. For example, if laser light illuminates a warm cloud of atoms from all directions and the laser's frequency is tuned below an atomic resonance, the atoms will be cooled. This common type of laser cooling relies on the Doppler effect where individual atoms will preferentially absorb laser light from the direction opposite to the atom's motion. The absorbed light is re-emitted by the atom in a random direction. After repeated emission and absorption of light the net effect on the cloud of atoms is that they will expand more slowly. The slower expansion reflects a decrease in the velocity distribution of the atoms, which corresponds to a lower temperature and therefore the atoms have been cooled. For an ensemble of particles, their thermodynamic temperature is proportional to the variance in their velocity, therefore the lower the distribution of velocities, the lower temperature of the particles.

<span class="mw-page-title-main">Optical tweezers</span> Scientific instruments

Optical tweezers are scientific instruments that use a highly focused laser beam to hold and move microscopic and sub-microscopic objects like atoms, nanoparticles and droplets, in a manner similar to tweezers. If the object is held in air or vacuum without additional support, it can be called optical levitation.

<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.

<span class="mw-page-title-main">Quantum Zeno effect</span> Quantum measurement phenomenon

The quantum Zeno effect is a feature of quantum-mechanical systems allowing a particle's time evolution to be slowed down by measuring it frequently enough with respect to some chosen measurement setting.

In physics, a Tonks–Girardeau gas is a Bose gas in which the repulsive interactions between bosonic particles confined to one dimension dominate the system's physics. It is named after physicists Marvin D. Girardeau and Lewi Tonks. It is not a Bose–Einstein condensate as it does not demonstrate any of the necessary characteristics, such as off-diagonal long-range order or a unitary two-body correlation function, even in a thermodynamic limit and as such cannot be described by a macroscopically occupied orbital in the Gross–Pitaevskii formulation.

<span class="mw-page-title-main">Electromagnetically induced transparency</span>

Electromagnetically induced transparency (EIT) is a coherent optical nonlinearity which renders a medium transparent within a narrow spectral range around an absorption line. Extreme dispersion is also created within this transparency "window" which leads to "slow light", described below. It is in essence a quantum interference effect that permits the propagation of light through an otherwise opaque atomic medium.

<span class="mw-page-title-main">Trapped-ion quantum computer</span> Proposed quantum computer implementation

A trapped-ion quantum computer is one proposed approach to a large-scale quantum computer. Ions, or charged atomic particles, can be confined and suspended in free space using electromagnetic fields. Qubits are stored in stable electronic states of each ion, and quantum information can be transferred through the collective quantized motion of the ions in a shared trap. Lasers are applied to induce coupling between the qubit states or coupling between the internal qubit states and the external motional states.

In condensed matter physics, Anderson localization 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.

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 in 1963 in the context of granular superconductors. 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.

In condensed matter physics, an ultracold atom is an atom with a temperature near absolute zero. At such temperatures, an atom's quantum-mechanical properties become important.

In spectroscopy, the Autler–Townes effect, is a dynamical Stark effect corresponding to the case when an oscillating electric field is tuned in resonance to the transition frequency of a given spectral line, and resulting in a change of the shape of the absorption/emission spectra of that spectral line. The AC Stark effect was discovered in 1955 by American physicists Stanley Autler and Charles Townes.

In atomic physics, Raman cooling is a sub-recoil cooling technique that allows the cooling of atoms using optical methods below the limitations of Doppler cooling, Doppler cooling being limited by the recoil energy of a photon given to an atom. This scheme can be performed in simple optical molasses or in molasses where an optical lattice has been superimposed, which are called respectively free space Raman cooling and Raman sideband cooling. Both techniques make use of Raman scattering of laser light by the atoms.

<span class="mw-page-title-main">Time crystal</span> Structure that repeats in time; a novel type or phase of non-equilibrium matter

In condensed matter physics, a time crystal is a quantum system of particles whose lowest-energy state is one in which the particles are in repetitive motion. The system cannot lose energy to the environment and come to rest because it is already in its quantum ground state. Time crystals were first proposed theoretically by Frank Wilczek in 2012 as a time-based analogue to common crystals – whereas the atoms in crystals are arranged periodically in space, the atoms in a time crystal are arranged periodically in both space and time. Several different groups have demonstrated matter with stable periodic evolution in systems that are periodically driven. In terms of practical use, time crystals may one day be used as quantum computer memory.

<span class="mw-page-title-main">Quantum simulator</span> Simulators of quantum mechanical systems

Quantum simulators permit the study of a quantum system in a programmable fashion. In this instance, simulators are special purpose devices designed to provide insight about specific physics problems. Quantum simulators may be contrasted with generally programmable "digital" quantum computers, which would be capable of solving a wider class of quantum problems.

<span class="mw-page-title-main">Tilman Esslinger</span> German physicist

Tilman Esslinger is a German experimental physicist. He is Professor at ETH Zurich, Switzerland, and works in the field of ultracold quantum gases and optical lattices.

Bose–Einstein condensation can occur in quasiparticles, particles that are effective descriptions of collective excitations in materials. Some have integer spins and can be expected to obey Bose–Einstein statistics like traditional particles. Conditions for condensation of various quasiparticles have been predicted and observed. The topic continues to be an active field of study.

In theoretical physics, the curvature renormalization group (CRG) method is an analytical approach to determine the phase boundaries and the critical behavior of topological systems. Topological phases are phases of matter that appear in certain quantum mechanical systems at zero temperature because of a robust degeneracy in the ground-state wave function. They are called topological because they can be described by different (discrete) values of a nonlocal topological invariant. This is to contrast with non-topological phases of matter that can be described by different values of a local order parameter. States with different values of the topological invariant cannot change into each other without a phase transition. The topological invariant is constructed from a curvature function that can be calculated from the bulk Hamiltonian of the system. At the phase transition, the curvature function diverges, and the topological invariant correspondingly jumps abruptly from one value to another. The CRG method works by detecting the divergence in the curvature function, and thus determining the boundaries between different topological phases. Furthermore, from the divergence of the curvature function, it extracts scaling laws that describe the critical behavior, i.e. how different quantities behave as the topological phase transition is approached. The CRG method has been successfully applied to a variety of static, periodically driven, weakly and strongly interacting systems to classify the nature of the corresponding topological phase transitions.

<span class="mw-page-title-main">Monika Aidelsburger</span> German quantum physicist

Monika Aidelsburger is a German quantum physicist, Professor and Group Leader at the Ludwig Maximilian University of Munich. Her research considers quantum simulation and ultra cold atomic gases trapped in optical lattices. In 2021, she was awarded both the Alfried-Krupp-Förderpreis and Klung Wilhelmy Science Award.

John Morrissey Doyle is an American physicist working in the field of Atomic, Molecular, and Optical (AMO) physics and Precision Particle Physics. He is the Henry B. Silsbee Professor of Physics, Director of the Japanese Undergraduate Research Exchange Program (JUREP), Co-Director of the Harvard Quantum Initiative as well as Co-director of the Ph.D. Program in Quantum Science and Engineering at Harvard University.

References

  1. 1 2 3 4 5 6 Grimm, Rudolf; Weidemüller, Matthias; Ovchinnikov, Yurii B. (2000), "Optical Dipole Traps for Neutral Atoms", Advances In Atomic, Molecular, and Optical Physics, Elsevier, pp. 95–170, arXiv: physics/9902072 , doi:10.1016/s1049-250x(08)60186-x, ISBN   978-0-12-003842-8, S2CID   16499267 , retrieved 2020-12-17
  2. 1 2 Bloch, Immanuel (October 2005). "Ultracold quantum gases in optical lattices". Nature Physics. 1 (1): 23–30. Bibcode:2005NatPh...1...23B. doi:10.1038/nphys138. S2CID   28043590.
  3. Gebhard, Florian (1997). The Mott metal-insulator transition models and methods. Berlin [etc.]: Springer. ISBN   978-3-540-61481-4.
  4. 1 2 Greiner, Markus; Mandel, Olaf; Esslinger, Tilman; Hänsch, Theodor W.; Bloch, Immanuel (January 3, 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.
  5. Koetsier, Arnaud; Duine, R. A.; Bloch, Immanuel; Stoof, H. T. C. (2008). "Achieving the Néel state in an optical lattice". Phys. Rev. A. 77 (2): 023623. arXiv: 0711.3425 . Bibcode:2008PhRvA..77b3623K. doi:10.1103/PhysRevA.77.023623. S2CID   118519083.
  6. Letokhov, V.S. (May 1968). "Narrowing of the Doppler Width in a Standing Wave" (PDF). Journal of Experimental and Theoretical Physics. 7: 272.
  7. Fallani, Leonardo; Fort, Chiara; Lye, Jessica; Inguscio, Massimo (May 2005). "Bose-Einstein condensate in an optical lattice with tunable spacing: transport and static properties". Optics Express. 13 (11): 4303–4313. arXiv: cond-mat/0505029 . Bibcode:2005OExpr..13.4303F. doi:10.1364/OPEX.13.004303. PMID   19495345. S2CID   27181534.
  8. Huckans, J. H. (December 2006). "Optical Lattices and Quantum Degenerate Rb-87 in Reduced Dimensions". University of Maryland Doctoral Dissertation.
  9. Li, T. C.; Kelkar,H.; Medellin, D.; Raizen, M. G. (April 3, 2008). "Real-time control of the periodicity of a standing wave: an optical accordion". Optics Express. 16 (8): 5465–5470. arXiv: 0803.2733 . Bibcode:2008OExpr..16.5465L. doi:10.1364/OE.16.005465. PMID   18542649. S2CID   11082498.
  10. 1 2 Bakr, Waseem S.; Gillen, Jonathon I.; Peng, Amy; Fölling, Simon; Greiner, Markus (2009-11-05). "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. ISSN   0028-0836. PMID   19890326. S2CID   4419426.
  11. Haller, Elmar; Hudson, James; Kelly, Andrew; Cotta, Dylan A.; Peaudecerf, Bruno; Bruce, Graham D.; Kuhr, Stefan (2015-09-01). "Single-atom imaging of fermions in a quantum-gas microscope". Nature Physics. 11 (9): 738–742. arXiv: 1503.02005 . Bibcode:2015NatPh..11..738H. doi:10.1038/nphys3403. hdl: 10023/8011 . ISSN   1745-2473. S2CID   51991496.
  12. Wei, Chenan; Sedrakyan, Tigran (2021-01-29). "Optical lattice platform for the Sachdev-Ye-Kitaev model". Phys. Rev. A. 103 (1): 013323. arXiv: 2005.07640 . Bibcode:2021PhRvA.103a3323W. doi:10.1103/PhysRevA.103.013323. S2CID   234363891.
  13. Derevianko, Andrei; Katori, Hidetoshi (3 May 2011). "Colloquium : Physics of optical lattice clocks". Reviews of Modern Physics. 83 (2): 331–347. arXiv: 1011.4622 . Bibcode:2011RvMP...83..331D. doi:10.1103/RevModPhys.83.331. S2CID   29455812.
  14. "Ye lab". Ye lab.
  15. Brennen, Gavin K.; Caves, Carlton; Jessen, Poul S.; Deutsch, Ivan H. (1999). "Quantum logic gates in optical lattices". Phys. Rev. Lett. 82 (5): 1060–1063. arXiv: quant-ph/9806021 . Bibcode:1999PhRvL..82.1060B. doi:10.1103/PhysRevLett.82.1060. S2CID   15297433.
  16. Yang, Bing; Sun, Hui; Hunag, Chun-Jiong; Wang, Han-Yi; Deng, Youjin; Dai, Han-Ning; Yuan, Zhen-Sheng; Pan, Jian-Wei (2020). "Cooling and entangling ultracold atoms in optical lattices". Science. 369 (6503): 550–553. arXiv: 1901.01146 . Bibcode:2020Sci...369..550Y. doi:10.1126/science.aaz6801. PMID   32554628. S2CID   219901015.
  17. Weidner, C. A.; Anderson, Dana Z. (27 June 2018). "Experimental Demonstration of Shaken-Lattice Interferometry". Physical Review Letters. 120 (26): 263201. arXiv: 1801.09277 . doi:10.1103/PhysRevLett.120.263201. PMID   30004774. S2CID   51625118.
  18. MacDonald, M. P.; Spalding, G. C.; Dholakia, K. (November 27, 2003). "Microfluidic sorting in an optical lattice". Nature. 426 (6965): 421–424. Bibcode:2003Natur.426..421M. doi:10.1038/nature02144. PMID   14647376. S2CID   4424652.