Application of Symmetry-Adapted Atomic Amplitudes
Abstract
:1. Introduction
2. Impact of Symmetry-Adapted Basis States
2.1. First Steps towards a General-Purpose Code
2.2. The Grasp Code: Successes, Capabilities and Limitations
2.3. Recent Demands on Relativistic Atomic Theory
- Since most atomic behavior is qualitatively well understood (and is often readily distinguishable from each other) within the atomic shell model and its underlying electron configurations, this “shell-model view” need to be recognized and, possibly, re-adjusted by the user during various steps of (complex) computations. This need arises especially for processes with free electrons in either the initial and/or final (scattering) state of the atoms. Apart from the set of reference configurations and from specifying the virtual excitation in the construction of restricted active spaces (RAS), the set-up of atomic computations should therefore enable one to readily reconstruct the underlying shell occupation at quite different level of the computations. The reconstruction and analysis of the important shells during the modeling then help determine and adapt the subspaces that are (to be) taken into account into the quantum representation of the atomic states. In fact, the simple control of the computations by means of selecting and discarding (groups of) electron configurations can hardly be overrated for atoms with multiple inner-shell holes.
- A more flexible treatment of the electron continuum will enable one to model resonances embedded into the continuum, as well as the ionization, recombination and decay dynamics of free atoms and ions. It will help also to incorporate the continuum (interchannel) interactions in the construction of scattering states [49] or to construct approximate atomic Green functions [50], cf. Section 3.1.
- The increasing complexity of atomic computations suggests to perform many, if not most, of its steps automatically, based on a set of well-chosen default parameters. Apart from the incorporation of relativistic corrections, this refers not only to the generation of the (mean-field) SCF but also to calculating the symmetry-adapted amplitudes (2)–(3) themselves, if the computational model and approximations can be expressed concisely enough. The simple use and overwriting of defaults values also determines the rate with which complex computations can be realized.
- Quite different observations are typically recorded by different spectroscopic communities owing to the—individual or coincident—measurement of photons, electrons and/or ions. The large number of experimental setups and scenarios then require a proper classification of atomic properties and processes as well as a language, close to the underlying physics, in order to avoid duplication and inhomogeneity of code, or the implementation of overly specific features.
- Recent advancements in short- and strong-field physics make it possible today to explore the electron dynamics of atoms under extreme conditions, and which clearly differ from standard (spectroscopic) measurements. Apart from a remarkable increase in the intensity of light pulses by six orders of magnitude or even more, the time structure of these pulses can be currently controlled down to a few cycles and help studying non-linear photoprocesses [51]. Moreover, the interaction of such pulses with atoms and ions often enforces (relativistic) theory to deal with time- and spatially-structured vector potentials , well beyond the typical electric-dipole approximation or the incorporation of additional multipoles. Still, these interactions can be expressed by symmetry-adapted many-electron amplitudes [52] and should therefore be handled by modern codes.
3. Access to and Application of Symmetry-Adapted Amplitudes
3.1. Towards Current Fields of Research and Applications
- (a)
- Excitation schemes based on the atomic shell model: Many atomic computations start from a given set of reference (electron) configurations in order to generate relevant subspaces of the many-electron Hilbert space. This is accomplished by specifying virtual excitations of (so-called) active electrons, while the selection of these subspaces certainly require a good physical understanding, there explicit setup is usually performed automatically. To support well-chosen subspaces for different applications, a proper notion need to be developed and implemented to readily manipulate ‘lists’ of electron configurations and configuration schemes. For the dielectronic recombination (DR) of a multiply-charged ion, for example, all configurations with a single (sub-) valence excitation as well as the capture of an additional electron into any valence shell, taken from a given list of such shells, contribute to some resonance strength and, hence, to the DR plasma rate coefficients, provided that the energy difference of the doubly-excited resonance and the initial level is positive, . For rate coefficients, in particular, the setup of the computations can be simplified considerably if such an excite-by-capture scheme is established and readily applied to the given set of reference configurations. The associated list of configurations is then generated automatically. There are many other applications of atomic structure theory which can be made much simpler by establishing the notion of proper excitation and de-excitation schemes, and which are all based on the shell model.
- (b)
- Coupling of atoms to the free-electron continuum: If atoms or ions are resonantly excited into the continuum of the next higher charge state, they are inherently coupled to this “infinity of states” owing to the (non-local) electron–electron interaction. In atomic theory, this coupling is often expressed by means of photo- and autoionization amplitudes, and with the electron–photon and electron–electron interaction matrix elements from above. Hereby, different ionization channels need to be distinguished due to the coupling of the —partial waves of—free electron with energy to the bound state of the remaining ion: as well as due to the selection rules of the considered process. Any successful implementation of ionization (and recombination) processes therefore needs a straight and quick access to these many-electron amplitudes, including one or several free electron waves .
- (c)
- Modeling of atomic cascades: Such cascades typically arise when inner-shell electrons are excited, or ionized, and subsequently lead to the emission of two or more electrons and/or photons. Indeed, atomic cascades are quite ubiquitous in nature and have therefore been explored for different scenarios, from precision measurements to the modeling of astrophysical spectra, and up to the search for exotic particles. Until the present, however, no quantitative analysis of such decay cascades appear often to be feasible owing to the large, or even huge, number of decay pathways which the atoms can undertake. Since most, if not all, of these cascades can normally be traced back to a proper combination of different symmetry-adapted amplitudes, they are suitable for general-purpose codes if the computation of these amplitudes is separated from the subsequent simulation of the associated ion, photon or electron distribution, or any other wanted information. In order to deal with different excitation and decay scenarios, Ref. [57] suggests a number of (cascade) schemes and approaches that are clearly discernible with regard to their complexity and computational costs. Typical schemes refer to the decay of inner-shell holes, including the prior excitation or ionization of electrons [58], the radiative or dielectronic capture of electrons [59], or the formation and photoemission from hollow ions. Many observations in astrophysics will be understood better, or still at all, if the interplay between the radiative and nonradiative decay pattern can be modeled with sufficient detail.
- (d)
- Request of atomic Green functions: In atomic and many-particle physics, Green functions often occur as propagators to formally represent the (integration over the) complete spectrum of the underlying Hamiltonian. Since such an integration over the complete spectrum is crucial also to many second- and higher-order perturbation processes, approximate atomic Green functions need to be developed and accessible rather similar to how the ASF can be applied.A simple access to approximate Green functions requires however a decomposition into building blocks that are suitable for atomic structure theory [50]. By making use of the rotational symmetry and parity of the ASF , each Green function can be split into separate channels (continua) of well-defined symmetry , quite similar to the one-electron Coulomb–Green function [60]. Within a finite basis, these channels then simply refer to a set of atomic levels , all of the same total symmetry . Using this notation, an approximate atomic Green function is given by an array (list) of k such channels and is formally written asWhile, formally, each Green function is built from an infinite number of such channels, and including both parities , only a few of these channels (continua) are in practice relevant for any non-linear interaction process, either because of the symmetry of the underlying interaction operators or due to further insights into the physics.
- (e)
- Nonlinear atomic processes: The explicit use of the many-electron interaction amplitudes from above also facilitates the implementation of nonlinear, e.g., second- and higher-order processes, once an appropriate Green function has been constructed for some given property or process of interest. Well-known second-order processes are the two-photon absorption and emission, the resonant and two-photon ionization, the radiative and double Auger emission of atoms, or the (single-photon) double ionization, to just recall a few of them. For each of these processes, different selection rules apply and help to restrict the number of Green function channels with total symmetry that need to be generated. In these computations, the summation (integration) remains the most challenging part of all numerical computations owing to number of terms in the representation of the Green function (5). Further difficulties may arise from the free-free matrix elements that occur in all ionization and capture processes. However, the recent interest and observation of multi-photon or multi-electron processes make the calculation of such nonlinear processes by general-purpose codes such as Grasp for sure highly desirable.
- (f)
- Different representation of atomic states: In the MCDHF method, the representation of the ASF is fully determined by choosing the energy functional of the SCF and the CSF basis in terms of virtual excitations with regard to a set of (bound) reference configurations. These atomic states can then be directly applied also to evaluate the interaction amplitudes and, hence, the observables of interest. Other representations are obtained by adopting the functional of the underlying mean field or by treating certain classes of excitation perturbatively. Again, many of these possible extensions of atomic theory are formulated in terms of the electron–electron interaction amplitudes from above. Moreover, rather similar amplitudes arise if the dominant QED corrections are to be taken into account by an single-electron model Hamiltonian [61]Apart from such ‘traditional’ many-body techniques, the qubit representation of the atomic Hamiltonian and the use of quantum hardware has attracted much recent interest. Different transformations, such as the Jordan–Wigner or Bravyi–Kitaev transformation, aim to map the electronic structure of indistinguishable fermions upon distinguishable qubits and to use algorithms that are exponentially faster than the best-known classical algorithms. Up to the present, however, no efficient algorithm is known to solve electronic-structure problems in a fully general form [63]. It will therefore be beneficial to explore how the symmetry-adapted amplitudes from above can be transformed and evaluated by means of qubit Hamiltonians and quantum circuits.
3.2. A Descriptive Language for Atomic Computations
- (1)
- Provide an intuitive user interface in which all computations can be expressed quite similar to the user’s research work;
- (2)
- Supply features for dealing with general open-shell configurations and applications in atomic physics and elsewhere; such features refer for instance to a self-acting generation of (lists of) configurations or Green function channels by just selecting suitable classes of virtual excitations;
- (3)
- Help redefine the physical units to the needs of the user, both at input and output time;
- (4)
- Give access to different models and approximations as well as the decomposition of a given task into well-designed steps, rather similar to writing pseudo-code;
- (5)
- Resort to default values whenever feasible; indeed, all parameters, which can be derived from other input or which is irrelevant for a given computation, should be based on default values, though with keeping the user informed. On the other hand, it must be quite easy to overwrite these defaults for selected applications;
- (6)
- Enable the user to redefine physical constants or default settings;
- (7)
- Support a transparent data flow with and within the program, independent of the shell structure of the atoms or any particular application;
- (8)
- Help visualize large data sets as often required for the analysis of results and for debugging;
- (9)
- Support the interactive use (test) of individual data types, functions or single lines of the code.
3.3. Impact of Proper Data Types
3.4. Jena Atomic Calculator
4. Summary and Conclusions
Funding
Acknowledgments
Conflicts of Interest
References
- Swirles, B. The relativistic self-consistent field. Proc. Roy. Soc. Lond. 1935, 152, 625. [Google Scholar]
- Grant, I.P. Relativistic self-consistent fields. Proc. Roy. Soc. A 1961, 262, 555. [Google Scholar] [CrossRef]
- Grant, I.P.; Burke, V.M. The effect of relativity on atomic wavefunctions. Proc. Phys. Soc. 1967, 90, 297. [Google Scholar]
- Desclaux, J.P. A multiconfiguration relativistic Dirac-Fock program. Comp. Phys. Commun. 1975, 9, 31. [Google Scholar] [CrossRef]
- Grant, I.P.; McKenzie, B.J.; Norrington, P.H.; Mayers, D.F.; Pyper, N.C. An atomic multiconfiguration Dirac-Fock package. Comp. Phys. Commun. 1980, 21, 207. [Google Scholar] [CrossRef]
- Parpia, F.A.; Fischer, C.F.; Grant, I.P. GRASP92: A package for large-scale relativistic atomic structure calculations. Comp. Phys. Commun. 1996, 94, 249. [Google Scholar] [CrossRef]
- Froese Fischer, C.; Gaigalas, G.; Jönsson, P.; Bieroń, J. GRASP2018—A Fortran 95 version of the General Relativistic Atomic Structure Package. Comp. Phys. Commun. 2019, 13, 520. [Google Scholar] [CrossRef]
- Judd, R. Operator Techniques in Atomic Spectroscopy; McGraw-Hill: New York, NY, USA, 1963. [Google Scholar]
- Johnson, W.R. Atomic Structure Theory: Lectures on Atomic Physics; Springer: Berlin/Heidelberg, Germany, 2007. [Google Scholar]
- Grant, I.P. Relativistic Quantum Theory of Atoms and Molecules: Theory and Computation; Springer: Berlin/Heidelberg, Germany, 2007. [Google Scholar]
- Indelicato, P. Available online: http://dirac.spectro.jussieu.fr/mcdf/mcdf_code/mcdfgme_status.html (accessed on 10 August 2022).
- Fritzsche, S. A toolbox for studying the properties of open-shell atoms and ions. J. Electron. Spectrosc. Relat. Phenom. 2001, 114–116, 1155. [Google Scholar] [CrossRef]
- Gaigalas, G.; Zalandauskas, T.; Fritzsche, S. Spectroscopic LSJ notation for atomic levels as obtained from relativistic calculations. Comput. Phys. Commun. 2004, 157, 239. [Google Scholar] [CrossRef] [Green Version]
- Fritzsche, S. A fresh computational approach to atomic structures, processes and cascades. Comp. Phys. Commun. 2019, 240, 1. [Google Scholar] [CrossRef]
- Grant, I.P. Relativistic calculation of atomic structures. Adv. Phys. 1970, 19, 747. [Google Scholar] [CrossRef]
- McKenzie, B.J.; Grant, I.P.; Norrington, P.H. A program to calculate transverse Breit and QED corrections to energy levels in a MCDF environment. Comp. Phys. Commun. 1980, 21, 233. [Google Scholar] [CrossRef]
- Brink, M.; Satchler, G.R. Angular Momentum, 2nd ed.; Clarendon Press: Oxford, UK, 1968. [Google Scholar]
- de-Shalit, A.; Feshbach, H. Theoretical Nuclear Physics: Nuclear Structure; Wiley: Hoboken, NJ, USA, 1974. [Google Scholar]
- Blatt, M.; Weisskopf, V.F. Theoretical Nuclear Physics; Springer: New York, NY, USA, 1979. [Google Scholar]
- Froese Fischer, C.; Godefroid, M.; Brage, T.; Jönsson, P.; Gaigalas, G. Advanced multiconfiguration methods for complex atoms: I. Energies and wave functions. J. Phys. B 2016, 49, 182004. [Google Scholar] [CrossRef] [Green Version]
- Froese Fischer, C.; Tachiev, G.; Irimia, A. Relativistic energy levels, lifetimes, and transition probabilities for the sodium-like to argon-like sequences. Atom. Data Nucl. Data Tabl. 2006, 92, 607. [Google Scholar] [CrossRef]
- Dong, C.Z.; Fritzsche, S.; Fricke, B.; Sepp, W.-D. Branching ratios and lifetimes of the low–lying levels of Fe X. Mon. Notes R. Astr. Soc. 1999, 307, 809. [Google Scholar] [CrossRef] [Green Version]
- Bieroń, J.; Froese Fischer, C.; Grant, I.P. Large-scale multiconfigurational Dirac-Fock calculations of the hyperfine-structure constants and determination of the nuclear quadrupole moment of 49Ti. Phys. Rev. A 1999, 59, 4295. [Google Scholar] [CrossRef]
- Bieroń, J.; Jönsson, P.; Froese Fischer, C. Effects of electron correlation, relativity, and nuclear structure on hyperfine constants of Be+ and F6+. Phys. Rev. A 1999, 60, 3547. [Google Scholar] [CrossRef]
- Fritzsche, S.; Zschornack, G.; Musiol, G.; Soff, G. Interchannel interactions in highly-energetic radiationless transitions of neonlike ions. Phys. Rev. A 1991, 44, 388. [Google Scholar] [CrossRef]
- Fritzsche, S.; Fricke, B.; Sepp, W.D. Reduced L1 level-width and Coster-Kronig yields by relaxation and continuum interactions in atomic zinc. Phys. Rev. A 1992, 45, 1465. [Google Scholar] [CrossRef] [Green Version]
- Sienkiewicz, J.E.; Fritzsche, S.; Grant, I.P. Relativistic configuration interaction approach to the elastic low-energy scattering of electrons from atoms. J. Phys. B 1995, 28, L633. [Google Scholar] [CrossRef]
- Grant, I.P. Relativistic Effects in Atoms and Molecules. In Methods in Computational Chemistry; Wilson, S., Ed.; Plenum: New York, NY, USA, 1988; Volume 2, pp. 1–38. [Google Scholar]
- Grant, I.P. A general program to calculate angular momentum coefficients in relativistic atomic structure. Comp. Phys. Commun. 1973, 4, 263. [Google Scholar] [CrossRef]
- Gaigalas, G.; Fritzsche, S.; Grant, I.P. Calculation of pure angular coefficients in jj-coupling. Comp. Phys. Commun. 2001, 139, 263. [Google Scholar] [CrossRef]
- Gaigalas, G.; Fritzsche, S. Angular coefficients for symmetry-adapted configuration states in jj-coupling. Comput. Phys. Commun. 2021, 267, 108086. [Google Scholar] [CrossRef]
- Filippin, L.; Bieroń, J.; Gaigalas, G.; Godefroid, M.; Jönsson, P. Multiconfiguration calculations of electronic isotope-shift factors in Zn I. Phys. Rev. A 2017, 96, 042502. [Google Scholar] [CrossRef] [Green Version]
- Eronen, T.; Fritzsche, S.; Geldhof, S.; Kelly, S.; Moore, I.D.; Pohjalainen, I.; Reponen, M.; Rinta-Antila, S.; Voss, A. Isotope shifts from collinear laser spectroscopy of doubly-charged yttrium isotopes. Phys. Rev. A 2018, 97, 042504. [Google Scholar]
- Fischer, C.F.; Gaigalas, G. The Effect of correlation on spectra of the lanthanides: Pr3+. Atoms 2018, 6, 8. [Google Scholar] [CrossRef] [Green Version]
- Perger, W.F.; Halabuka, Z.; Trautmann, D. Continuum wavefunction solver for GRASP. Comp. Phys. Commun. 1993, 76, 250. [Google Scholar] [CrossRef]
- Zhang, H.L.; Sampson, D.H. Relativistic distorted-wave collision strengths and oscillator strengths for all possible n = 2 − n = 3 transitions in b-like ions. Atom. Data Nucl. Data Tabl. 1994, 58, 255. [Google Scholar] [CrossRef]
- Fritzsche, S. The Ratip program for relativistic calculations of atomic transition, ionization and recombination properties. Comp. Phys. Commun. 2012, 183, 1525. [Google Scholar] [CrossRef]
- Saha, B.; Fritzsche, S. Influence of dense plasma on the low-lying transitions in Be-like ions: Relativistic multiconfiguration Dirac–Fock calculation. J. Phys. B 2007, 40, 259. [Google Scholar] [CrossRef]
- Norrington, P.H.; Grant, I.P. Low-energy electron scattering by Fe XXIII and Fe VII using the Dirac R-matrix method. J. Phys. B 1987, 20, 4869. [Google Scholar] [CrossRef]
- Olsen, J.; Godefroid, M.; Jönsson, P.; Målmqvist, P.A.; Froese Fischer, C. Transition probability calculations for atoms using nonorthogonal orbitals. Phys. Rev. E 1995, 52, 4499. [Google Scholar] [CrossRef] [PubMed]
- Fritzsche, S.; Grant, I.P. A program for the complete expansion of jj-coupled symmetry functions into Slater determinants. Comp. Phys. Commun. 1995, 92, 111. [Google Scholar] [CrossRef]
- Jönsson, P.; Froese Fischer, C. Multiconfiguration Dirac-Fock calculations of the 2s2 1S0 − 2s2p 3P1 intercombination transition in C III. Phys. Rev. A 1998, 57, 4967. [Google Scholar] [CrossRef]
- Yordanov, D.T.; Balabanski, D.L.; Bieroń, J.; Bissell, M.L.; Blaum, K.; Budinčević, I.; Fritzsche, S.; Frömmgen, N.; Georgiev, G.; Geppert, C.; et al. Spins, Electromagnetic moments and isomers of 107–129Cd. Phys. Rev. Lett. 2013, 110, 192501. [Google Scholar] [CrossRef] [Green Version]
- Raeder, S.; Ackermann, D.; Backe, H.; Block, M.; Cheal, B.; Chhetri, P.; Düllmann, C.E.; Van Duppen, P.; Even, J.; Ferrer, R.; et al. Nuclear properties of nobelium isotopes from laser spectroscopy. Phys. Rev. Lett. 2018, 120, 232503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
- Eliav, E.; Fritzsche, S.; Kaldor, U. Electronic structure theory of the superheavy elements. Nucl. Phys. 2015, A944, 518. [Google Scholar] [CrossRef]
- Bieroń, J.; Fischer, C.F.; Fritzsche, S.; Gaigalas, G.; Grant, I.P.; Indelicato, P.; Jönsson, P.; Pyykkö, P. Ab initio MCDHF calculations of electron-nucleus interactions. Phys. Scr. 2015, 90, 054011. [Google Scholar] [CrossRef] [Green Version]
- Bieroń, J.; Filippin, L.; Gaigalas, G.; Godefroid, M.; Jönsson, P.; Pyykko, P. Ab initio calculations of the hyperfine structure of zinc and evaluation of the nuclear quadrupole moment Q (Zn-67). Phys. Rev. A 2018, 97, 062505. [Google Scholar] [CrossRef] [Green Version]
- Sato, T.K.; Asai, M.; Tsukada, K.; Kaneya, Y.; Toyoshima, A.; Makii, H.; Mitsukai, A.; Nagame, Y.; Osa, A.; Toyoshima, A.; et al. First ionization potentials of Fm, Md, No, and Lr: Verification of filling-up of 5f electrons and confirmation of the actinide series. J. Amer. Chem. Soc. 2018, 140, 14609. [Google Scholar] [CrossRef] [Green Version]
- Åberg, T.; Howat, G. Theory of the Auger Effect. In Corpuscles and Radiation in Matter I; Encyclopedia of Physics Vol. XXXI; Mehlhorn, W., Ed.; Springer: Berlin/Heidelberg, Germany, 1982; p. 469. [Google Scholar]
- Fritzsche, S.; Surzhykov, A. Approximate atomic Green functions. Molecules 2021, 26, 2660. [Google Scholar] [CrossRef] [PubMed]
- Feldhaus, J. FLASH—The first soft X-ray free electron laser (FEL) user facility. J. Phys. B 2010, 43, 194002. [Google Scholar] [CrossRef]
- Fritzsche, S.; Böning, B. Lorentz-force shifts in strong-field ionization with mid-infrared laser fields. Phys. Rev. Res. 2022, 4, 033031. [Google Scholar] [CrossRef]
- Charlwood, F.C.; Billowes, J.; Campbell, P.; Cheal, B.; Eronen, T.; Forest, D.H.; Fritzsche, S.; Honma, M.; Jokinen, A.; Moore, I.D.; et al. Ground state properties of manganese isotopes across the N = 28 shell closure. Phys. Lett. B 2010, 690, 346. [Google Scholar] [CrossRef]
- Cheal, B.; Cocolios, T.E.; Fritzsche, S. Laser spectroscopy of radioactive isotopes: Role and limitations of accurate isotope-shift calculations. Phys. Rev. A 2012, 86, 042501. [Google Scholar] [CrossRef]
- Ferrer, R.; Barzakh, A.; Bastin, B.; Beerwerth, R.; Block, M.; Creemers, P.; Grawe, H.; de Groote, R.; Delahaye, P.; Fléchard, X.; et al. Towards high-resolution laser ionization spectroscopy of the heaviest elements in supersonic gas jet expansion. Nature Commun. 2017, 8, 14520. [Google Scholar] [CrossRef] [Green Version]
- Surzhykov, A.; Jentschura, U.D.; Stöhlker, T.; Gumberidze, A.; Fritzsche, S. Alignment of heavy few–electron ions following excitation by relativistic Coulomb collisions. Phys. Rev. A 2008, 77, 042722. [Google Scholar] [CrossRef]
- Fritzsche, S.; Palmeri, P.; Schippers, S. Atomic cascade computations. Symmetry 2021, 13, 520. [Google Scholar] [CrossRef]
- Schippers, S.; Martins, M.; Beerwerth, R.; Bari, S.; Holste, K.; Schubert, K.; Viefhaus, J.; Savin, D.W.; Fritzsche, S.; Müller, A.; et al. Near L-edge single and multiple photoionization of singly charged iron ions. Astrophys. J. 2017, 849, 5. [Google Scholar] [CrossRef] [Green Version]
- Fritzsche, S. Dielectronic recombination strengths and plasma rate coefficients of multiply-charged ions. Astron. Astrophys. 2021, 656, A163. [Google Scholar] [CrossRef]
- Wong, M.K.F.; Ye, E.H.Y. The Dirac Coulomb Green’s function and its application to relativistic Rayleigh scattering. J. Math. Phys. 1985, 26, 1701. [Google Scholar] [CrossRef]
- Shabaev, V.M.; Tupitsyn, I.I.; Yerokhin, V.A. Model operator approach to the Lamb shift calculations in relativistic many-electron atoms. Phys. Rev. A 2013, 88, 012513. [Google Scholar] [CrossRef] [Green Version]
- Indelicato, P.; Santos, J.P.; Boucard, S.; Desclaux, J.-P. QED and relativistic corrections in superheavy elements. Eur. Phys. J. D 2007, 45, 155. [Google Scholar] [CrossRef]
- Kandala, A.; Mezzacapo, A.; Temme, K.; Takita, M.; Brink, M.; Chow, J.M.; Gambetta, J.M. Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets. Nature 2017, 549, 242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
- Julia 1.8 Documentation. Available online: https://docs.julialang.org/ (accessed on 10 August 2022).
- Bezanson, J.; Chen, J.; Chung, B.; Karpinski, S.; Shah, V.B.; Vitek, J.; Zoubritzky, L. Julia: Dynamism and performance reconciled by design. Proc. ACM Program. Lang. 2018, 2, 120. [Google Scholar] [CrossRef]
- Kwong, T. Hands-On Design Patterns and Best Practices with Julia; Packt Publishing: Birmingham, UK, 2020. [Google Scholar]
- Fritzsche, S. JAC: User Guide, Compendium & Theoretical Background. Available online: https://github.com/OpenJAC/JAC.jl (accessed on 10 August 2022).
Struct | Brief Explanation |
---|---|
AbstractEeInteraction | Abstract type to distinguish between different electron–electron interaction operators; it comprises the concrete (singleton) types BreitInteraction, CoulombInteraction, CoulombBreit. |
AbstractExcitationScheme | Abstract type to support different excitation schemes, such as DeExciteSingleElectron, DeExciteTwoElectrons, ExciteByCapture, and several others. |
AbstractGreenApproach | Defines an abstract type for approximating a many-electron Green function expansion, and which comprises the concrete (singleton) types SingleCSFwithoutCI and CoreSpaceCI. |
AbstractScField | Abstract type for dealing with different self-consistent-field (SCF) potentials. |
AsfSettings | Settings to control the SCF and CI calculations for a given multiplet of ASF. |
Atomic.Computation | An atomic computation of one or several multiplets, including the SCF and CI calculations, as well as of selected properties or processes. |
Basis | Relativistic atomic basis, including the full specification of the configuration space and radial orbitals. |
Configuration | Non-relativistic electron configuration in terms of its shell occupation. |
ConfigurationR | Relativistic electron configuration in terms of its subshell occupation. |
EmMultipole | A multipole component of the electro-magnetic field, such as E1, M1, E2, … and as specified by its parity and multipolarity. |
GreenChannel | A single approximate Green function channel with well-defined symmetry . |
Level | Atomic level in terms of its quantum numbers, symmetry, energy and its (possibly full) representation. |
LevelSelection | List of levels that is specified by either the level numbers and/or level symmetries. |
LevelSymmetry | specifies the total angular momentum and parity of a particular level. |
LSjjSettings | Settings to control the transformation of the selected many-electron levels. |
Multiplet | An ordered list of atomic levels, often associated with one or several configurations. |
Nuclear.Model | A nuclear model of an atom to keep all nuclear parameters together. |
Orbital | Relativistic radial orbital function that appears as ‘buildung block’ in order to define the many-electron CSF; its is typically given on a (radial) grid and comprises a large and small component. |
Radial.Grid | Radial grid to represent the (radial) orbitals and to perform all radial integrations. |
Radial.Potential | Radial potential function. |
Shell | Non-relativistic shell, such as . |
Subshell | Relativistic subshell, such as |
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. |
© 2022 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Fritzsche, S. Application of Symmetry-Adapted Atomic Amplitudes. Atoms 2022, 10, 127. https://doi.org/10.3390/atoms10040127
Fritzsche S. Application of Symmetry-Adapted Atomic Amplitudes. Atoms. 2022; 10(4):127. https://doi.org/10.3390/atoms10040127
Chicago/Turabian StyleFritzsche, Stephan. 2022. "Application of Symmetry-Adapted Atomic Amplitudes" Atoms 10, no. 4: 127. https://doi.org/10.3390/atoms10040127
APA StyleFritzsche, S. (2022). Application of Symmetry-Adapted Atomic Amplitudes. Atoms, 10(4), 127. https://doi.org/10.3390/atoms10040127