Jump to content

Rydberg atom: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Ojovan (talk | contribs)
Line 86: Line 86:
Rydberg atoms form commonly in [[Plasma (physics)|plasma]]s due to the recombination of electrons and positive ions; low energy recombination results in fairly stable Rydberg atoms, while recombination of electrons and positive ions with high [[kinetic energy]] often form [[Ionization potential|autoionising]] Rydberg states. Rydberg atoms’ large sizes and susceptibility to perturbation and ionisation by electric and magnetic fields, are an important factor determining the properties of plasmas<ref name="Plasma">{{cite journal |author=G. Vitrant et al |title=Rydberg to plasma evolution in a dense gas of very excited atoms |journal=Journal of Physics B: Atomic and Molecular Physics |volume=15 |issue=2 |pages=L49–L55 |year=1982 |url=http://www.iop.org/EJ/article/0022-3700/15/2/004/jbv15i2pL49.pdf?request-id=22fbca2a-9b9b-4d2e-83da-3c0732670172 |doi= 10.1088/0022-3700/15/2/004|doi_brokendate=}}</ref>.
Rydberg atoms form commonly in [[Plasma (physics)|plasma]]s due to the recombination of electrons and positive ions; low energy recombination results in fairly stable Rydberg atoms, while recombination of electrons and positive ions with high [[kinetic energy]] often form [[Ionization potential|autoionising]] Rydberg states. Rydberg atoms’ large sizes and susceptibility to perturbation and ionisation by electric and magnetic fields, are an important factor determining the properties of plasmas<ref name="Plasma">{{cite journal |author=G. Vitrant et al |title=Rydberg to plasma evolution in a dense gas of very excited atoms |journal=Journal of Physics B: Atomic and Molecular Physics |volume=15 |issue=2 |pages=L49–L55 |year=1982 |url=http://www.iop.org/EJ/article/0022-3700/15/2/004/jbv15i2pL49.pdf?request-id=22fbca2a-9b9b-4d2e-83da-3c0732670172 |doi= 10.1088/0022-3700/15/2/004|doi_brokendate=}}</ref>.


Condensation of Rydberg atoms forms [[Rydberg matter]]<ref name="Matter">{{cite journal |author=E. A. Manykin et al |title=Rydberg matter: properties and decay |journal=Proceedings of the SPIE |volume=6181 |issue= |pages=618105:1–9 |year=2006 |url=http://spiedl.aip.org/getabs/servlet/GetabsServlet?prog=normal&id=PSISDG006181000001618105000001&idtype=cvips&gifs=yes |doi=10.1117/12.675004 |doi_brokendate=}}</ref> most often observed in form of long-lived clusters.
Condensation of Rydberg atoms forms [[Rydberg matter]] most often observed in form of long-lived clusters. The de-excitation is significantly impeded in Rydberg matter by exchange-correlation effects in the non-uniform electron liquid formed on condensation by the collective valence electrons, which causes extended lifetime of clusters<ref name="Matter">{{cite journal |author=E. A. Manykin et al |title=Rydberg matter: properties and decay |journal=Proceedings of the SPIE |volume=6181 |issue= |pages=618105:1–9 |year=2006 |url=http://spiedl.aip.org/getabs/servlet/GetabsServlet?prog=normal&id=PSISDG006181000001618105000001&idtype=cvips&gifs=yes |doi=10.1117/12.675004 |doi_brokendate=}}</ref>.


=== Rydberg atoms in astrophysics ===
=== Rydberg atoms in astrophysics ===

Revision as of 11:40, 8 May 2009

A Rydberg atom is an excited atom with one or more electrons that have a very high principal quantum number[1]. These atoms have a number of peculiar properties including an exaggerated response to electric and magnetic fields[2], long decay periods and electron wavefunctions that approximate, under some conditions, classical orbits of electrons about the nuclei. The core electrons shield the outer electron from the electric field of the nucleus such that, from a distance, the electric potential looks identical to that experienced by the electron in a hydrogen atom[3].

In spite of its shortcomings, the Bohr model of the atom is useful in explaining these properties. Classically an electron in a circular orbit of radius r, about a nucleus of charge Ze, obeys Newton's second law:

where k = 1/(4πε0).

Orbital momentum is quantised in units of ħ:

.

Combining these two equations leads to Bohr's expression for the orbital radius in terms of the principal quantum number, n:

It is now apparent why Rydberg atoms have such peculiar properties; the radius of the orbit scales as n2 (the n = 137 state of hydrogen has an atomic radius ~1 µm) and the geometric cross-section as n4. Thus Rydberg atoms are extremely large with loosely bound valence electrons, easily perturbed or ionized by collisions or external fields.

The classical view of Rydberg atoms

Figure 1. A comparison of the potential in a hydrogen atom with that in a Rydberg state of a different atom. A large core polarizability has been used in order to make the effect clear. The black curve is the Coulombic 1/r potential of the hydrogen atom while the dashed red curve includes the 1/r4 term due to polarization of the ion core.

An atom in a Rydberg state has a valence electron in a large orbit far from the ion core; in such an orbit the outermost electron feels an almost hydrogenic, Coulomb potential, VC from a compact ion core consisting of a nucleus with Z protons and the lower electron shells filled with Z-1 electrons. An electron in the spherically symmetric Coulomb potential has potential energy:

.

The similarity of the effective potential ‘seen’ by the outer electron to the hydrogen potential is a defining characteristic of Rydberg states and explains why the electron wavefunctions approximate to classical orbits in the limit of the correspondence principle[4]. There are three notable exceptions that can be characterized by the additional term added to the potential energy:

  • An atom may have two (or more) electrons in highly excited states with comparable orbital radii. In this case the electron-electron interaction gives rise to a significant deviation from the hydrogen potential[5]. For an atom in a multiple Rydberg state, the additional term, Uee, includes a summation of each pair of highly excited electrons:
.
  • If the valence electron has very low angular momentum (interpreted classically as an extremely eccentric elliptical orbit) then it may pass close enough to polarise the ion core, giving rise to a 1/r4 core polarization term in the potential[6]:
,
where αd is the dipole polarizability. Figure 1 shows how the polarization term modifies the potential close to the nucleus.
  • If the outer electron penetrates the inner electron shells, it will 'see' more of the charge of the nucleus and hence experience a greater force.

The quantum mechanical picture

Figure 2. Semiclassical orbits for n=5 with all allowed values of orbital angular momentum. The black spot denotes the position of the atomic nucleus.

Quantum mechanically a state with high n refers to an atom in which the valence electron(s) have been excited into a formerly unpopulated electron orbital with higher energy and lower binding energy. In hydrogen the binding energy is given by:

,

where Ry = 13.6 eV is the Rydberg constant. The low binding energy at high values of n explains why Rydberg states are susceptible to ionization.

Additional terms in the potential energy expression for a Rydberg state, on top of the hydrogenic Coulomb potential energy require the introduction of a quantum defect, δl, into the expression for the binding energy:

.

The long lifetimes of Rydberg states with high orbital angular momentum can be explained in terms of the overlapping of wavefunctions. The wavefunction of an electron in a high l state (high angular momentum, 'circular orbit') has very little overlap with the wavefunctions of the inner electrons and hence remains relatively unperturbed.

The three exceptions to the definition of a Rydberg atom as an atom with a hydrogenic potential, have an alternative, quantum mechanical description that can be characterized by the additional term(s) in the atomic Hamiltonian:

  • If a second electron is excited into a state ni, energetically close to the state of the outer electron no, then its wavefunction become almost as large as the first (a double Rydberg state). This occurs as ni approaches no and leads to a condition where the size of the two electron’s orbits are related[5]; a condition sometimes referred to as radial correlation[1]. An electron-electron repulsion term must be included in the atomic Hamiltonian.
  • Polarization of the ion core produces an anisotropic potential that causes an angular correlation[1][7] between the motions of the two outermost electrons. This can be thought of as a tidal locking effect due to a non-spherically symmetric potential. A core polarization term must be included in the atomic Hamiltonian.
  • The wavefunction of the outer electron in states with low orbital angular momentum l, is periodically localised within the shells of inner electrons and interacts with the full charge of the nucleus[8]. Figure 2 shows a semi-classical interpretation of angular momentum states in an electron orbital, illustrating that low-l states pass closer to the nucleus potentially penetrating the ion core. A core penetration term must be added to the atomic Hamiltonian.

What makes Rydberg atoms worth studying?

Experimental atomic physics

Although the interesting nature of Rydberg atoms has been understood since the development of the Bohr model in 1913, for a long time there was no efficient method for producing large, single-state, mono-energetic populations. There were a number of interesting experiments done in the 1950s and 1960s using electron impact excitation[9]:

and charge exchange excitation[10]:

but these techniques both produce populations with a very broad spread of energies.

The arrival of tunable dye lasers in the 1970s allowed a much greater level of control over populations of excited atoms. In optical excitation the incident photon is absorbed by the atom, thus absolutely specifying the energy of the state produced.

The precise control over atomic states allowed by optical excitation has played a significant role in advancing our understanding of atomic Rydberg states.

Investigating diamagnetic effects

The large sizes and low binding energies of Rydberg atoms lead to a high magnetic susceptibility, Χ. As diamagnetic effects scale with the area of the orbit and the area is proportional to the radius squared (A α n4), effects impossible to detect in ground state atoms become obvious in Rydberg atoms, which demonstrate very large diamagnetic shifts[11].

Rydberg atoms in plasmas

Rydberg atoms form commonly in plasmas due to the recombination of electrons and positive ions; low energy recombination results in fairly stable Rydberg atoms, while recombination of electrons and positive ions with high kinetic energy often form autoionising Rydberg states. Rydberg atoms’ large sizes and susceptibility to perturbation and ionisation by electric and magnetic fields, are an important factor determining the properties of plasmas[12].

Condensation of Rydberg atoms forms Rydberg matter most often observed in form of long-lived clusters. The de-excitation is significantly impeded in Rydberg matter by exchange-correlation effects in the non-uniform electron liquid formed on condensation by the collective valence electrons, which causes extended lifetime of clusters[13].

Rydberg atoms in astrophysics

In the time between the early absorption spectroscopy experiments and the arrival of tunable lasers, interest in Rydberg atoms was kept alive by the realisation that they are common in interstellar space, and as such are an important radiation source for astronomers.

The density within interstellar gas clouds is typically many orders of magnitude lower than the best laboratory vacuums attainable on Earth, allowing Rydberg atoms to persist for long periods of time without being ionised by collisions or electric and magnetic fields. As a result of this longevity and the abundance of hydrogen it is particularly common for astronomers to observe radiation from the heavens at a frequency of 2.4 GHz, now known to correspond to the hydrogen n = 109 to n = 108 transition[14]. Such a highly excited hydrogen atom on Earth would be ionised almost immediately as the binding energy would be significantly below thermal energies.

Strongly interacting systems

Due to their large size, Rydberg atoms can exhibit very large electric dipole moments. Calculations using perturbation theory show that this results in strong interactions between two close Rydberg atoms. Coherent control of these interactions combined with the their relatively long lifetime makes them a suitable candidate to realize a quantum computer [15]. As of March 2009, a two-qubit gate has not been achieved experimentally; however, observations of collective excitations or conditional dynamics have been reported, both between two individual atoms [16] [17] and in mesoscopic samples [18]. Strongly interacting Rydberg atoms also feature quantum critical behavior [19], which makes them interesting to study on their own.

Classical simulation of a Rydberg atom

Figure 3. Stark - Coulomb potential for a Rydberg atom in a static electric field. An electron in such a potential feels a torque that can change its angular momentum.
Figure 4. Trajectory of the electron in a hydrogen atom in an electric field E = -3 x 106 V/m in the x-direction. Note that classically all values of angular momentum are allowed; figure 2 shows the particular orbits associated with quantum mechanically allowed values. See the animation.

A simple 1/r potential results in a closed Keplerian elliptical orbit. In the presence of an external electric field Rydberg atoms can obtain very large electric dipole moments making them extremely susceptible to perturbation by the field. Figure 3 shows how application of an external electric field (known in atomic physics as a Stark field) changes the geometry of the potential, dramatically changeing the behaviour of the electron. A Coulombic potential does not apply any torque as the force is always antiparallel to the position vector (always pointing along a line running between the electron and the nucleus):

,
.

With the application of a static electric field, the electron feels a continuously changing torque. The resulting trajectory becomes progressively more distorted over time, eventually going through the full range of angular momentum from L = LMAX, to a straight line L=0, to the initial orbit in the opposite sense L = -LMAX[20].

The time period of the oscillation in angular momentum (the time to complete the trajectory in figure 4), almost exactly matches the quantum mechanically predicted period for the wavefunction to return to its initial state, demonstrating the classical nature of the Rydberg atom.

See also

References

  1. ^ a b c Gallagher, Thomas F. (1994). Rydberg Atoms. Cambridge: Cambridge University Press. ISBN 0521021669. {{cite book}}: Cite has empty unknown parameter: |coauthors= (help)
  2. ^ Metcalf Research Group (2004-11-08). "Rydberg Atom Optics". Stoney Brook University. Retrieved 2008-07-30.
  3. ^ Nolan, James (2005-05-31). "Rydberg Atoms and the Quantum Defect". Davidson College. Retrieved 2008-07-30. {{cite web}}: Cite has empty unknown parameter: |coauthors= (help)
  4. ^ T. P. Hezel; et al. (1992). "Classical view of the properties of Rydberg atoms: Application of the correspondence principle". American Journal of Physics. 60 (4): 329–335. doi:10.1119/1.16876. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help); Explicit use of et al. in: |author= (help)
  5. ^ a b I. K. Dmitrieva and G. I. Plindov (1993). "Energies of Doubly Excited Sates. The Double Rydberg Formula". Journal of Applied Spectroscopy. 59 (1–2): 466–470. doi:10.1007/BF00663353. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help)
  6. ^ L. Neale and M. Wilson (1995). "Core Polarization in Kr VIII". Physical Review A. 51 (5): 4272–4275. doi:10.1103/PhysRevA.51.4272. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help)
  7. ^ T. A. Heim and A. R. P. Rau (1995). "Excitation of high-lying pair-Rydberg states". Journal of Physics B: Atomic, Molecular and Optical Physics. 28: 5309–5315. doi:10.1088/0953-4075/28/24/015. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help)
  8. ^ C. E. Theodosiou. "Evaluation of penetration effects in high-l Rydberg states". {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help); Cite journal requires |journal= (help)
  9. ^ J. Olmsted (1967). "Excitation of nitrogen triplet states by electron impact". Radiation Research. 31 (2): 191–200. doi:10.2307/3572319. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help)
  10. ^ M. Haugh; et al. (1966). "Electronic excitation accompanying charge exchange". Journal of Chemical Physics. 44 (2): 837–839. doi:10.1063/1.1726773. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help); Explicit use of et al. in: |author= (help)
  11. ^ J. Neukammer; et al. (1984). "Diamagnetic shift and singlet-triplet mixing of 6snp Yb Rydberg states with large radial extent". Physical Review A. 30 (2): 1142–1144. doi:10.1103/PhysRevA.30.1142. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help); Explicit use of et al. in: |author= (help)
  12. ^ G. Vitrant; et al. (1982). "Rydberg to plasma evolution in a dense gas of very excited atoms" (PDF). Journal of Physics B: Atomic and Molecular Physics. 15 (2): L49–L55. doi:10.1088/0022-3700/15/2/004. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help); Explicit use of et al. in: |author= (help)
  13. ^ E. A. Manykin; et al. (2006). "Rydberg matter: properties and decay". Proceedings of the SPIE. 6181: 618105:1–9. doi:10.1117/12.675004. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help); Explicit use of et al. in: |author= (help)
  14. ^ J. Poutanen (2007-03-22). "Gasdynamics and Interstellar Medium" (PDF). University of Oulu. Retrieved 2008-07-31. {{cite web}}: line feed character in |title= at position 16 (help)
  15. ^ D. Jaksch; et al. (2000). "Fast Quantum Gates for Neutral Atoms". Physical Review Letters. 85 (10): 2208–2211. doi:10.1103/PhysRevLett.85.2208. arXiv:quant-ph/0004038. {{cite journal}}: Explicit use of et al. in: |author= (help)
  16. ^ A. Gaëtan; et al. (2009). "Observation of collective excitation of two individual atoms in the Rydberg blockade regime". Nature Physics. 5 (2): 115–118. doi:10.1038/nphys1183. arXiv:0810.2960. {{cite journal}}: Explicit use of et al. in: |author= (help)
  17. ^ E. Urban; et al. (2009). "Observation of Rydberg blockade between two atoms". Nature Physics. 5 (2): 110–114. doi:10.1038/nphys1178. arXiv:0805.0758. {{cite journal}}: Explicit use of et al. in: |author= (help)
  18. ^ R. Heidemann; et al. (2007). "Evidence for Coherent Collective Rydberg Excitation in the Strong Blockade Regime". Physical Review Letters. 99 (16): 163601. doi:10.1103/PhysRevLett.99.163601. arXiv:quant-ph/0701120. {{cite journal}}: Explicit use of et al. in: |author= (help)
  19. ^ H. Weimer; et al. (2008). "Quantum Critical Behavior in Strongly Interacting Rydberg Gases". Physical Review Letters. 101 (25): 250601. doi:10.1103/PhysRevLett.101.250601. arXiv:0806.3754. {{cite journal}}: Explicit use of et al. in: |author= (help)
  20. ^ T. P. Hezel; et al. (1992). "Classical view of the Stark effect in hydrogen atoms". American Journal of Physics. 60 (4): 324–328. doi:10.1119/1.16875. {{cite journal}}: Cite has empty unknown parameter: |doi_brokendate= (help); Explicit use of et al. in: |author= (help)