The bare quantum states for an atom in a strong, monochromatic AC field on the left are defined by a photon number and whether the atom is in its ground (gnd) or excited (ex) state. The AC stark effect creates "dressed" states and subsequently three three different transition energies. Mollow triplet, is seen in the fluorescence spectrum to the right.[1]


The AC Stark effect (also known as the dynamic(al) Stark effect or Autler-Towns effect) refers to the shift of energy levels for an atom or molecule in an alternating electric field. It is the AC equivalent of the Stark effect which splits the spectral lines of atoms and molecules in a constant electric field. Compared to its DC counterpart though, the fields in the AC effect are typically much larger, and the effects are much harder to predict.[2]

While generally referring to atomic spectral shifts due to AC fields at any (single) frequency, the effect is dramatically enhanced when the field is tuned to the frequency of a natural two level transition. [3] In this case, the alternating field has the effect of splitting the two bare transition states into doublets or "dressed states" that are separated by the Rabi frequency. [4] This is commonly achieved by a laser tuned to (or near) the desired transition.

This splitting results in a Rabi cycle or Rabi oscillation between bare states which are no longer energy eigenstates of the atom-field Hamiltonian.[5] The resulting fluorescence spectrum of an atom is known as a Mollow triplet. The AC stark splitting is integral to several other phenomena in quantum optics, such as Electromagnetically induced transparency and Sisyphus cooling. Vacuum Rabi oscillations have also been described as manifestation of the AC Stark effect from atomic coupling to the vacuum field. [4]

History

edit

Often also referred to as the Autler-Townes effect, the AC stark effect was discovered in 1955 by the eponymous American physicists Stanley Autler and Charles Townes while at Columbia University and Lincoln Labs at the Massachusetts Institute of Technology. Before the availability of shorter wavelength coherent sources (namely, lasers), the AC Stark effect was observed with radio frequency sources. Autler and Townes' original observation of the effect used a radio frequency source tuned to 12.78 and 38.28 MHz, corresponding to the separation between two doublet microwave absorption lines of OCS. [6]

The notion of quasi-energy in treating the general AC Stark effect was later developed by Nikishov and Ritis in 1964 and onward [7][8] [9]. This more general method of approaching the problem developed into the celebrated "dressed atom" model describing the interaction between lasers and atoms[5].

Prior to the 1970s there were various conflicting predictions concerning the fluorescence spectra of atoms due to the AC Stark effect at optical frequencies. In 1974 the observation of Mollow triplets verified the form of the AC Stark effect using visible light[3].

General Semiclassical Approach

edit

In a semiclassical model where the electromagnetic field is treated classically, a system of charges in a monochromatic electromagnetic field has a Hamiltonian that can be written as:

 

where  ,  ,   and   are respectively the position, momentum, mass, and charge of the  -th particle, and   is the speed of light. The vector potential of the field,  , satisfies

 .

The Hamiltonian is thus also periodic:

 

Now, the Schrödinger equation, under a periodic Hamiltonian is a linear homogeneous differential equation with periodic coefficients,

 

where   here represents all coördinates. Floquet's theorem guarantees that the solutions to an equation of this form can be written as

 

Here,   is the "bare" energy for no coupling to the electromagnetic field, and   has the same time-periodicity as the Hamiltonian,

 

or

 

with   the angular frequency of the field.

Because of it's periodicity, it is often further useful to expand   in a Fourier series, obtaining

 

or

 


where   is the frequency of the laser field.


The solution for the joint particle-field system is therefore a linear combination of stationary states of energy  , which is known as a quasi-energy state and the new set of energies are called the spectrum of quasi-harmonics.[9]


Unlike the DC Stark effect, where perturbation theory is useful in a general case of atoms with infinite bound states, obtaining even a limited spectrum of shifted energies for the AC Stark effect is difficult in all but simple models, although calculations for systems such as the hydrogen atom have been done. [10]

Examples

edit

General expressions for AC Stark shifts must usually be calculated numerically and tend to provide little insight[2]. However, there are important individual examples of the effect that are informative.

Two level atom dressing

edit

An atom driven by an electric field with frequency   close to an atomic transition (consider detuning  ) can be approximated as a two level quantum system since the off resonance states have low occupation probability.[4] The Hamiltonian can be divided into the bare atom term plus a term for the interaction with the field as:

 

In an appropriate rotating frame, and making the rotating wave approximation,   reduces to

 

Where   is the Rabi frequency, and   are the strongly coupled bare atom states. The energy eigenvalues are

 

and for small detuning,

 

The eigenstates of the atom-field system or dressed states are dubbed   and  .


The result of the AC field on the atom is thus to shift the strongly coupled bare atom energy eigenstates states into two states   and   which are now separated by  . Evidence of this shift is apparent in the atom's absorption spectrum, which shows two peaks around the bare transition frequency, seperated by   (Autler-Townes splitting). The modified absorption spectrum can be obtained by a pump-probe experiment, wherein a strong pump laser drives the bare transition while a weaker probe laser sweeps for a second transition between a third atomic state and the dressed states.


Another consequence of the AC Stark splitting here is the appearance of Mollow triplets, a triple peaked fluorescence profile. Historically an important confirmation of Rabi flopping, they were first predicted by Mollow in 1969 [11] and confirmed in the 1970s experimentally. [4]

Adiabatic elimination

edit

In quantum system with three (or more) states, where a transition from one level,   to another   can be driven by an AC field, but   only decays to states other than  , the dissipative influence of the spontaneous decay can be eliminated. This is achieved by increasing the AC Stark shift on   through large detuning and raising intensity of the driving field. Adiabatic elimination has been used to create comparatively stable effective two level systems in Rydberg atoms, which are of interest for qubit manipulations in quantum computing.[12] [13] [14]

Electromagnetically induced transparency

edit

Electromagnetically induced transparency (EIT), which gives some materials a small transparent area within an absorption line, can be thought of as a combination of Autler-Townes splitting and Fano interference, although the distinction may be difficult to determine experimentally. While both Autler-Towns splitting and EIT can produce a transparent window in an absorption band, EIT refers to a window that maintains transparency in a weak pump field, and thus requires Fano interference. Because Autler-Townes splitting will wash out Fano interference at stronger fields, a smooth transition between the two effects is evident in materials exhibiting EIT.[15]

See also

edit


References

edit
  1. ^ Santori, Charles (2009). "Quantum dots: Driven to perfection". Nature Physics. 5: 173–174. doi:10.1038/nphys1215. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  2. ^ a b Delone, Nikolai Borisovich, and Vladimir Pavlovich Krainov. "AC Stark shift of atomic energy levels." Physics-Uspekhi 42.7 (1999): 669-687.
  3. ^ a b Schuda, F., C. R. Stroud Jr, and M. Hercher. "Observation of the resonant Stark effect at optical frequencies." Journal of Physics B: Atomic and Molecular Physics 7.7 (1974): L198.
  4. ^ a b c d Fox, Mark. Quantum Optics: An Introduction: An Introduction. Vol. 15. Oxford university press, 2006.
  5. ^ a b Barnett, Stephen, and Paul M. Radmore. Methods in theoretical quantum optics. Vol. 15. Oxford University Press, 2002.
  6. ^ Autler, Stanley H., and C. H. Townes. "Stark effect in rapidly varying fields." Physical Review 100.2 (1955): 703.
  7. ^ Nikishov, A. I., and V. I. Ritus. QUANTUM PROCESSES IN THE FIELD OF A PLANE ELECTROMAGNETIC WAVE AND IN A CONSTANT FIELD. PART I. Lebedev Inst. of Physics, Moscow, 1964.
  8. ^ Ritus, V. I. "Shift and splitting of atomic energy levels by the field of an electromagnetic wave." Sov. Phys. JETP 24 (1967): 1041-1044.
  9. ^ a b Zel'Dovich, Ya B. "Scattering and emission of a quantum system in a strong electromagnetic wave." Physics-Uspekhi 16.3 (1973): 427-433.
  10. ^ Crance, Michèle. "Nonperturbative ac Stark shifts in hydrogen atoms." JOSA B 7.4 (1990): 449-455.
  11. ^ B.R. Mollow, Phys. Rev. 188, 1969
  12. ^ Brion, Etienne, Line Hjortshøj Pedersen, and Klaus Mølmer. "Adiabatic elimination in a lambda system." Journal of Physics A: Mathematical and Theoretical 40.5 (2007): 1033.
  13. ^ Radmore, P_M_, and P. L. Knight. "Population trapping and dispersion in a three-level system." Journal of Physics B: Atomic and Molecular Physics 15.4 (1982): 561.
  14. ^ Linskens, A. F., et al. "Two-photon Rabi oscillations." Physical Review A 54.6 (1996): 4854.
  15. ^ Anisimov, Petr M., Jonathan P. Dowling, and Barry C. Sanders. Autler-Townes Splitting vs. Electromagnetically Induced Transparency: Objective Criterion to Discern Between Them in any Experiment. No. arXiv: 1102.0546. 2011.