Next Article in Journal
Inclusion of Electron Interactions by Rate Equations in Chemical Models
Next Article in Special Issue
A Progress Report on Laser Resonance Chromatography
Previous Article in Journal
The Binary-Encounter-Bethe Model for Computation of Singly Differential Cross Sections Due to Electron-Impact Ionization
Previous Article in Special Issue
The NEXT Project: Towards Production and Investigation of Neutron-Rich Heavy Nuclides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Developments in the Production and Research of Actinide Elements

by
Mustapha Laatiaoui
1,2,3,* and
Sebastian Raeder
2,3,*
1
Department Chemie—Standort TRIGA, Johannes Gutenberg-Universität Mainz, 55099 Mainz, Germany
2
Helmholtz Institut Mainz, 55099 Mainz, Germany
3
GSI Helmholtzzentrum für Schwerionenforschung, 64291 Darmstadt, Germany
*
Authors to whom correspondence should be addressed.
Atoms 2022, 10(2), 61; https://doi.org/10.3390/atoms10020061
Submission received: 30 April 2022 / Revised: 27 May 2022 / Accepted: 1 June 2022 / Published: 8 June 2022
(This article belongs to the Special Issue Atomic Structure of the Heaviest Elements)

Abstract

:
This article briefly reviews topics related to actinide research discussed at the virtual workshop Atomic Structure of Actinides & Related Topics organized by the University of Mainz, the Helmholtz Institute Mainz, and the GSI Helmholtz Centre for Heavy Ion Research, Darmstadt, Germany, and held on the 26–28 May 2021. It includes references to recent theoretical and experimental work on atomic structure and related topics, such as element production, access to nuclear properties, trace analysis, and medical applications.

1. Introduction

The actinides refer to the series of elements at the bottom of the periodic table, which includes elements from actinium (Ac, Z = 89) to lawrencium (Lr, Z = 103). Apart from the current controversy about their positioning in the periodic table and about the composition of “Group 3” [1,2], the research on actinides is experiencing an upsurge with the development of sensitive and sophisticated spectroscopic techniques and regularly attracts attention in the scientific media. It is a multidisciplinary research area par excellence, encompassing experimental and theoretical atomic and nuclear physics, nuclear chemistry, and quantum chemistry, with links to astrophysics, nuclear medicine, nuclear forensics, trace analysis, and so on. A virtual workshop entitled Atomic Structure of Actinides & Related Topics was held on the 26–28 May 2021 and brought together many experts in the abovementioned areas to present and discuss their recent findings and outlooks. To commemorate this successful event, we have included the logo of the workshop in this article; see Figure 1. This current issue of the MDPI journal Atoms includes some of these scientific contributions for the same reason and in the hope of serving as a useful resource for future studies of the atomic structure of the heaviest elements. In this article we provide a brief overview of the topics presented at the workshop, focusing on aspects of atomic structure, without claiming completeness. For the various areas covered here, other review articles are available, to which we refer in the appropriate places.

2. Element Production

The lack of availability and the radioactive nature of most actinide elements render their experimental investigation challenging. The lighter actinides up to uranium (U, Z = 92) are still found in the earth’s crust today because they are either primordial, i.e., they were produced before the formation of the solar system and are not yet fully decayed due to their long half-lives, as is the case for 232Th ( T 1 / 2 = 1.405 × 10 10 a), and 238U ( T 1 / 2 = 4.468 × 10 9 a), or because they are continuously produced from the decay of the primordial ones [3]. The heavier actinides up to fermium can be produced in high flux reactors by breeding lighter actinide elements in successive chains of neutron capture reactions with subsequent β decays. Due to competing fission processes and the still-limited neutron flux, the quantities produced in such nuclear reactors become smaller and smaller with increasing atomic number. Although milligram quantities can still be obtained for californium (Cf, Z = 98), only microgram or even nanogram amounts are available (at most) for einsteinium (Es, Z = 99) and fermium (Fm, Z = 100), respectively [4,5]. The breeding chain ends as soon as the short-lived fission isotopes of the latter element are reached.
To produce elements beyond fermium, nuclear reaction mechanisms such as fusion evaporation reactions induced by accelerated heavy ion beams are exploited instead [6,7]. Often projectile currents in the range of one particle μA from large accelerator facilities are used in conjunction with thin-target production techniques. In this way, the heaviest actinides from mendelevium to lawrencium can be produced at rates of a few atoms per second at most, whereas heavier elements up to oganesson can be produced at much lower rates.
With multi-nucleon-transfer reactions at energies close or slightly above the Coulomb barrier, one may target the production of more exotic and rather neutron-rich isotopes of the heavy elements [8] at existing facilities such as the KISS facility at RIKEN, Japan [9], and the IGISOL facility at JYFL in Jyväskylä, Finland [10], or at upcoming ones such as the N = 126 -factory at ANL in Argonne, USA [11], and the NEXT project at the AGOR cyclotron facility in Groningen, Netherlands [7].
Far away from earth-based accelerators, different scenarios are conceivable as to how these elements could have been or are being produced in nature. Currently, neutron star mergers are considered to be the most promising candidates for the so-called rapid neutron capture process [3] for the production of such elements due to the enormous neutron fluxes involved. Detection would therefore require multimessenger astronomy and a detailed spectral analysis of the light coming from these distant sources. However, in order to interpret these processes in terms of element abundances and production yields through comparison with the latest astrophysical models, a profound knowledge of the spectral light emissions of the elements and thus their atomic structure is essential [12,13,14,15].

3. Atomic Structure Modeling

In the second half of the last century, many experimentalists and theorists devoted increasing attention to the study of the atomic structure of the actinides, which until then had been a poorly explored area. It has been recognized that relativistic effects play an important role in the theoretical description of these atomic species [16]. Therefore, relativistic modeling of such atoms is indispensable today. Nevertheless, their accurate description remains a challenge, especially because they exhibit an open 5 f shell, which provides space for a large number of valence electrons and thus access to numerous low-lying electron configurations. A prominent example of an actinide with a complex atomic structure is protactinium (Pa, Z = 91), for which a detailed analysis of its atomic level density already indicates quantum chaos behavior, as reported in [17,18]. Therefore, constraints on modeling these atomic systems are often direct implications of the greatly increased configuration space [19,20,21,22].
Meanwhile, various numerical approaches, such as multi-configuration Hartree–Dirac Fock (MCDHF), Fock-space coupled-cluster (FSCC), and configuration-interaction (CI), are constantly being developed to provide reliable predictions for ever more complex systems [22,23,24]. Only recently has progress been made in the ab initio framework, such that the atomic properties and spectra of actinium (Ac, Z = 89) [25], fermium (Fm, Z = 100) [26], mendelevium (Md, Z = 101) [27], lawrencium (Lr, Z = 103) [28,29,30], rutherfordium (Rf, Z = 104) [31,32], and dubnium (Db, Z = 105) [33] can be calculated with a relatively high degree of reliability.
These calculations guide and complement experimental studies of the atomic structure of the heavy actinides, as was accomplished several years ago for fermium [34,35] and nobelium (No, Z = 102) [36,37,38,39,40,41], and also predict the behavior of known superheavy elements that remain long out of the reach of experimental investigation, as is the case with oganesson (Og, Z = 118) [42,43,44]. In the recent years they have helped in pinning down the first ionization potential of the heaviest actinides that were measured utilizing surface ionization techniques [45,46] and in unveiling extraordinary possibilities for the realization of optical clocks based on highly charged actinide ions [47,48].

4. Experiments Targeting the Atomic Structure

Across the actinides, a wealth of information on their atomic structures can be obtained from databases such as NIST, albeit mainly for the lighter representatives of this series [49,50]. As the atomic number increases, the scarcity of information becomes apparent, not least because experimental investigations become increasingly complex and often tedious due to decreasing sample quantities. The methods used are often adapted to the element production process and optimized for studies on that specific element, and would usually need to be further developed to achieve higher sensitivity to cover other exotic or heavier radionuclides. We limit ourselves here to a few examples; for a detailed overview of some of these developments, the reader is referred to [51] and the references therein.
One example of contemporary experimental investigations of the heaviest actinides is the RADRIS technique. It is based on laser resonance ionization in a gas cell and was developed for atomic level searches on the element nobelium. The technique enabled the experimental observation and characterization of an optical transition of this element for the first time [41], which subsequently paved the way to extract the first ionization potential with high precision [52]. The RADRIS technique is constantly being optimized, with the focus now on spectroscopy of the next heavy element, lawrencium—the last member of the actinide series [53].
Another example is the novel method of laser resonance chromatography. Researchers have recently been proposed to extend laser spectroscopy to lawrencium and the refractory metals of the superheavy elements, enabling both atomic-level searches and subsequent high-resolution spectroscopy [54,55]. The method exploits optical pumping to change the transport properties of an ion as it drifts in helium gas, allowing optical resonances to be identified based on drift times alone, without using resonance ionization or fluorescence detection. A corresponding experimental setup is currently under construction [56]. Closely related to this, and due to the fact that transport properties are sensitive to electronic configurations, efforts are currently being devoted to making ion mobility spectrometry one of the useful tools for isobaric purification in actinide research [57].

5. Nuclear Properties

Another pertinent topic in this regard is the fact that optical spectroscopy can bridge atomic and nuclear physics. Because the interaction of the central nucleus with the atomic shell leads to subtle but measurable differences in the atomic line spectrum, hyperfine spectroscopy can be used to study nuclear properties such as nuclear electric and magnetic moments and changes in the mean square charge radii [58,59]. The question of which atomic transitions are most suitable for hyperfine spectroscopy depends essentially on the requirement that the sensitivity of the experimental method should not be greatly deteriorated and on the precision with which one can determine the nuclear parameters. Again, calculations based on atomic theory are often required, since some nuclear properties cannot be measured absolutely and—due to the scarcity of information—comparisons with isotopes already studied via other methods are only possible in particular cases.
Such studies are of paramount importance in the field of heavy elements. The fact that the actinides span from the closed spherical neutron shell with neutron number N = 126 to the deformed shell closure at N = 152 makes them an ideal experimental field for the study of shell effects and nuclear deformation, which are essential for understanding nuclear stability in the presence of the extreme repulsive Coulomb forces between the numerous protons in the nucleus.
Spectroscopically, interest in the study of actinides has increased in recent years due to the increasing sensitivity and selectivity of the applied techniques. Long-lived plutonium (Pu, Z = 94 ) isotopes were studied via laser spectroscopy in an atomic beam through resonance ionization spectroscopy and in a fast ion beam through collinear fluorescence techniques with high spectral resolution [60], and also for production in a buffer gas cell [61]. Actinium (Ac, Z = 89) was investigated via in-gas-jet, in-gas-cell, and hot cavity laser spectroscopy [62,63,64,65], whereas the latter technique could only recently be applied to californium (Cf, Z = 98 ) [66] and einsteinium (Es, Z = 99 ) isotopes [67].
For the study of even heavier elements, one may emphasize two examples. The RADRIS technique enabled experimental studies of 252–254No-isotopes, which, using atomic theory, allowed, for instance, the extraction of the charge radii changes [68]. In addition to level searches on lawrencium, RADRIS experiments are continuing, targeting heavy actinides to reach more exotic radionuclides with half-lives ranging from below one second up to more than one hour [53], given that the atomic transitions are known.
The other example is the in-gas-jet laser spectroscopy. This technique exploits laser resonance ionization in a low-temperature and low-density environment of a collimated gas jet emerging from a buffer gas stopping cell to enable high spectral resolution and maintain a high sensitivity at the same time [65,69,70,71]. Setups using his technique are planned for the S3 separator at GANIL, Caen [72,73,74]; the MARA separator at JYFL, Jyväskylä [75]; and at the SHIP separator at GSI, Darmstadt [76,77]. Since the extraction of sample radionuclides occurs relatively quickly, one may gain access to short-lived nuclear isomers such as the K-isomer of 254No. With the development of the in-gas-jet technique, the required laser systems are being further developed [78,79,80] to accommodate the needs of high-resolution spectroscopy.
Other research in relation to heavy elements includes the study of higher-order deformations of nuclei [81] and studies of radioactive molecules [82], which are motivated by the search for new particle physics beyond the standard model.
For more details on laser spectroscopy, we refer the reader to recent reviews [58,59] in a broader context and to [51] in the context of actinide research.

6. The Thorium-229 Nuclear Isomer

Another topic of particular interest for fundamental research is the low-lying nuclear isomer of thorium-229, the existence of which was demonstrated experimentally a few years ago [83]. The isomeric state is only 8.28 eV [84,85] above the ground state of the nucleus and thus within the VUV range of optical probing. Due to this singular property, several applications regarding a nuclear clock based on optical transitions of this isomer have been proposed [86], although several open questions concerning this isotope are still being investigated [87,88].
In many experiments the isomer is produced in the α-decay branch of 233U. A promising new approach is currently being pursued to prepare the isomer online at increased efficiency via the β decay of 229Ac [89]. In all experiments, however, the most important and urgent goal today is to determine the excitation energy with an accuracy that should allow much more precise spectroscopic studies in the future [90]. Since the energy of the nuclear isomer is in the range of atomic binding energies, internal conversion can dominate radiative decay, especially in a neutral atomic state [91,92]. This, along with the need for laser cooling in future applications, brings 2+ and 3+ cations of this isotope more into focus in this research field [93,94].

7. Trace Analysis and Medical Applications

As radionuclides of trans-uranium elements are artificially produced in nuclear reactors, their elemental and isotopic abundance in test samples depends on the fuel material and the neutron flux initially used. Therefore, quantitative analysis of these telltale isotopic abundances can be exploited to identify radioactive material and to determine the history of the sample for nuclear forensics, safeguards, and proliferation [95,96]. Some trace analysis techniques rely on laser resonance ionization or light absorption and therefore also require a thorough knowledge of atomic structure. For resonance ionization spectroscopy (RIS) of actinides, the development of efficient excitation and ionization schemes is ongoing [97,98,99,100], and recent reports attest to advances in the use of spatially resolved secondary neutral mass spectrometry combined with RIS on sample atoms [96,101,102].
Another technical application for actinide research arises from developments in targeted alpha-therapy, in which the locally released energy in alpha decay has been used for the treatment of cancer cells. For this medical application, alpha emitters such as 225Ac are being studied in detail, which in turn requires an understanding of atomic properties, production and purification, and the underlying chemical behavior for selective binding to cancer cells [103,104].

8. Summary

The actinides in general, and the heaviest of them in particular, are very limited in availability. Because of this limitation and the lack of applications for most of them, even their basic physical properties have been scarcely studied. The increasing interest in the above-mentioned research aspects related to the atomic structure of actinides and the availability of increasingly sensitive spectroscopy techniques has recently pushed actinide research forward. Experimental and theoretical groups are jointly working on gradually filling in gaps in our knowledge regarding the atomic and nuclear properties of these fascinating elements at the bottom of the periodic table, so more exciting results can be expected in the near future.

Author Contributions

All authors contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

M.L. acknowledges funding from the European Research Council (ERC) under the European Union’s Horizon 2020 Research and Innovation Programme (Grant Agreement No. 819957).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vernon, R.E. The location and composition of Group 3 of the periodic table. Found. Chem. 2021, 23, 155–197. [Google Scholar] [CrossRef]
  2. de Bettencourt-Dias, A. The Periodic Table and the f Elements. In Rare Earth Elements and Actinides: Progress in Computational Science Applications; ACS Publications: Washington, DC, USA, 2021; pp. 55–61. [Google Scholar]
  3. Cowan, J.J.; Sneden, C.; Lawler, J.E.; Aprahamian, A.; Wiescher, M.; Langanke, K.; Martínez-Pinedo, G.; Thielemann, F.K. Origin of the heaviest elements: The rapid neutron-capture process. Rev. Mod. Phys. 2021, 93, 015002. [Google Scholar] [CrossRef]
  4. Roberto, J.B.; Rykaczewski, K.P. Actinide Targets for the Synthesis of Superheavy Nuclei: Current Priorities and Future Opportunities. In Proceedings of the Fission and Properties of Neutron-Rich Nuclei Sixth International Conference on ICFN6, Sanibel Island, FL, USA, 6–12 November 2016; World Scientific: Singapore, 2017. [Google Scholar] [CrossRef]
  5. Robinson, S.M.; Benker, D.E.; Collins, E.D.; Ezold, J.G.; Garrison, J.R.; Hogle, S.L. Production of Cf-252 and other transplutonium isotopes at Oak Ridge National Laboratory. Radiochim. Acta 2020, 108, 737–746. [Google Scholar] [CrossRef]
  6. Moody, K.J. Synthesis of Superheavy Elements. In The Chemistry of Superheavy Elements; Springer: Berlin/Heidelberg, Germany, 2013; pp. 1–81. [Google Scholar] [CrossRef]
  7. Even, J.; Chen, X.; Soylu, A.; Fischer, P.; Karpov, A.; Saiko, V.; Saren, J.; Schlaich, M.; Schlathölter, T.; Schweikhard, L.; et al. The NEXT project: Towards production and investigation of neutron-rich heavy nuclides. Atoms 2022, 10, 59. [Google Scholar] [CrossRef]
  8. Münzenberg, G.; Devaraja, H.M.; Dickel, T.; Geissel, H.; Gupta, M.; Heinz, S.; Hofmann, S.; Plass, W.R.; Scheidenberger, C.; Winfield, J.S.; et al. SHE Research with Rare-Isotope Beams, Challenges and Perspectives, and the New Generation of SHE Factories. In New Horizons in Fundamental Physics; Schramm, S., Schäfer, M., Eds.; Springer International Publishing: Cham, Switzerland, 2017; pp. 81–90. [Google Scholar] [CrossRef]
  9. Miyatake, H. KISS project. AIP Conf. Proc. 2021, 2319, 080006. Available online: https://aip.scitation.org/doi/pdf/10.1063/5.0036990 (accessed on 5 February 2022). [CrossRef]
  10. Dickel, T.; Kankainen, A.; Spătaru, A.; Amanbayev, D.; Beliuskina, O.; Beck, S.; Constantin, P.; Benyamin, D.; Geissel, H.; Gröf, L.; et al. Multi-nucleon transfer reactions at ion catcher facilities—A new way to produce and study heavy neutron-rich nuclei. J. Phys. Conf. Ser. 2020, 1668, 012012. [Google Scholar] [CrossRef]
  11. Savard, G.; Brodeur, M.; Clark, J.; Knaack, R.; Valverde, A. The N = 126 factory: A new facility to produce very heavy neutron-rich isotopes. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2020, 463, 258–261. [Google Scholar] [CrossRef]
  12. Smartt, S.J.; Chen, T.W.; Jerkstrand, A.; Coughlin, M.; Kankare, E.; Sim, S.A.; Fraser, M.; Inserra, C.; Maguire, K.; Chambers, K.C.; et al. A kilonova as the electromagnetic counterpart to a gravitational-wave source. Nature 2017, 551, 75–79. [Google Scholar] [CrossRef]
  13. Fontes, C.J.; Fryer, C.L.; Hungerford, A.L.; Wollaeger, R.T.; Korobkin, O. A line-binned treatment of opacities for the spectra and light curves from neutron star mergers. Mon. Not. R. Astron. Soc. 2020, 493, 4143–4171. Available online: https://academic.oup.com/mnras/article-pdf/493/3/4143/32920641/staa485.pdf (accessed on 27 May 2022). [CrossRef] [Green Version]
  14. Tanaka, M.; Kato, D.; Gaigalas, G.; Kawaguchi, K. Systematic opacity calculations for kilonovae. Mon. Not. R. Astron. Soc. 2020, 496, 1369–1392. Available online: https://academic.oup.com/mnras/article-pdf/496/2/1369/33424936/staa1576.pdf (accessed on 27 May 2022). [CrossRef]
  15. Silva, R.F.; Sampaio, J.M.; Amaro, P.; Flörs, A.; Martínez-Pinedo, G.; Marques, J.P. Structure Calculations in Nd III and U III Relevant for Kilonovae Modelling. Atoms 2022, 10, 18. [Google Scholar] [CrossRef]
  16. Pyykkö, P. The RTAM electronic bibliography, version 17.0, on relativistic theory of atoms and molecules. J. Comput. Chem. 2013, 34, 2667. [Google Scholar] [CrossRef] [PubMed]
  17. Naubereit, P.; Studer, D.; Viatkina, A.V.; Buchleitner, A.; Dietz, B.; Flambaum, V.V.; Wendt, K. Intrinsic quantum chaos and spectral fluctuations within the protactinium atom. Phys. Rev. A 2018, 98, 022506. [Google Scholar] [CrossRef]
  18. Naubereit, P.; Gottwald, T.; Studer, D.; Wendt, K. Excited atomic energy levels in protactinium by resonance ionization spectroscopy. Phys. Rev. A 2018, 98, 022505. [Google Scholar] [CrossRef]
  19. Eliav, E.; Borschevsky, A.; Kaldor, U. Electronic Structure at the Edge of the Periodic Table. Nucl. Phys. News 2019, 29, 16–20. [Google Scholar] [CrossRef]
  20. Fritzsche, S. Symbolic Evaluation of Expressions from Racah’s Algebra. Symmetry 2021, 13, 1558. [Google Scholar] [CrossRef]
  21. Safronova, M.S.; Safronova, U.I.; Kozlov, M.G. Atomic properties of actinide ions with particle-hole configurations. Phys. Rev. A 2018, 97, 012511. [Google Scholar] [CrossRef] [Green Version]
  22. Fritzsche, S. Level Structure and Properties of Open f-Shell Elements. Atoms 2022, 10, 7. [Google Scholar] [CrossRef]
  23. Dzuba, V. Calculation of Polarizabilities for Atoms with Open Shells. Symmetry 2020, 12, 1950. [Google Scholar] [CrossRef]
  24. Kahl, E.; Berengut, J. Ambit: A programme for high-precision relativistic atomic structure calculations. Comput. Phys. Commun. 2019, 238, 232–243. [Google Scholar] [CrossRef] [Green Version]
  25. Dzuba, V.A.; Flambaum, V.V.; Roberts, B.M. Calculations of the atomic structure for the low-lying states of actinium. Phys. Rev. A 2019, 100, 022504. [Google Scholar] [CrossRef] [Green Version]
  26. Allehabi, S.O.; Li, J.; Dzuba, V.; Flambaum, V. Theoretical study of electronic structure of erbium and fermium. J. Quant. Spectrosc. Radiat. Transf. 2020, 253, 107137. [Google Scholar] [CrossRef]
  27. Li, J.; Dzuba, V. Theoretical study of the spectroscopic properties of mendelevium (Z = 101). J. Quant. Spectrosc. Radiat. Transf. 2020, 247, 106943. [Google Scholar] [CrossRef] [Green Version]
  28. Kahl, E.V.; Berengut, J.C.; Laatiaoui, M.; Eliav, E.; Borschevsky, A. High-precision ab initio calculations of the spectrum of Lr+. Phys. Rev. A 2019, 100, 062505. [Google Scholar] [CrossRef] [Green Version]
  29. Kahl, E.V.; Raeder, S.; Eliav, E.; Borschevsky, A.; Berengut, J.C. Ab initio calculations of the spectrum of lawrencium. Phys. Rev. A 2021, 104, 052810. [Google Scholar] [CrossRef]
  30. Ramanantoanina, H.; Borschevsky, A.; Block, M.; Laatiaoui, M. Electronic structure of Lr+ (Z = 103) from ab initio calculations. Atoms 2022, 10, 48. [Google Scholar] [CrossRef]
  31. Ramanantoanina, H.; Borschevsky, A.; Block, M.; Laatiaoui, M. Electronic structure of Rf+ (Z = 104) from ab initio calculations. Phys. Rev. A 2021, 104, 022813. [Google Scholar] [CrossRef]
  32. Allehabi, S.O.; Dzuba, V.A.; Flambaum, V.V. Theoretical study of the electronic structure of hafnium (Hf,Z = 72) and rutherfordium (Rf,Z = 104) atoms and their ions: Energy levels and hyperfine-structure constants. Phys. Rev. A 2021, 104, 052811. [Google Scholar] [CrossRef]
  33. Lackenby, B.G.C.; Dzuba, V.A.; Flambaum, V.V. Calculation of atomic spectra and transition amplitudes for the superheavy element Db (Z = 105). Phys. Rev. A 2018, 98, 022518. [Google Scholar] [CrossRef] [Green Version]
  34. Sewtz, M.; Backe, H.; Dretzke, A.; Kube, G.; Lauth, W.; Schwamb, P.; Eberhardt, K.; Grüning, C.; Thörle, P.; Trautmann, N.; et al. First Observation of Atomic Levels for the Element Fermium (Z = 100). Phys. Rev. Lett. 2003, 90, 163002. [Google Scholar] [CrossRef] [Green Version]
  35. Backe, H.; Dretzke, A.; Fritzsche, S.; Haire, R.G.; Kunz, P.; Lauth, W.; Sewtz, M.; Trautmann, N. Laser Spectroscopic Investigation of the Element Fermium (Z = 100). Hyperfine Interact. 2005, 162, 3–14. [Google Scholar] [CrossRef]
  36. Fritzsche, S. On the accuracy of valence–shell computations for heavy and super–heavy elements. Eur. Phys. J. D 2005, 33, 15–21. [Google Scholar] [CrossRef]
  37. Borschevsky, A.; Eliav, E.; Vilkas, M.J.; Ishikawa, Y.; Kaldor, U. Predicted spectrum of atomic nobelium. Phys. Rev. A 2007, 75, 042514. [Google Scholar] [CrossRef]
  38. Indelicato, P.; Santos, J.P.; Boucard, S.; Desclaux, J.P. QED and relativistic corrections in superheavy elements. Eur. Phys. J. D 2007, 45, 155–170. [Google Scholar] [CrossRef]
  39. Liu, Y.; Hutton, R.; Zou, Y. Atomic structure of the super-heavy element No I (Z = 102). Phys. Rev. A 2007, 76, 062503. [Google Scholar] [CrossRef]
  40. Dzuba, V.A.; Safronova, M.S.; Safronova, U.I. Atomic properties of superheavy elements No, Lr, and Rf. Phys. Rev. A 2014, 90, 012504. [Google Scholar] [CrossRef] [Green Version]
  41. Laatiaoui, M.; Lauth, W.; Backe, H.; Block, M.; Ackermann, D.; Cheal, B.; Chhetri, P.; Düllmann, C.E.; van Duppen, P.; Even, J.; et al. Atom-at-a-time laser resonance ionization spectroscopy of nobelium. Nature 2016, 538, 495–498. [Google Scholar] [CrossRef]
  42. Lackenby, B.G.C.; Dzuba, V.A.; Flambaum, V.V. Atomic structure calculations of superheavy noble element oganesson (Z = 118). Phys. Rev. A 2018, 98, 042512. [Google Scholar] [CrossRef] [Green Version]
  43. Guo, Y.; Pašteka, L.F.; Eliav, E.; Borschevsky, A. Ionization potentials and electron affinity of oganesson with relativistic coupled cluster method. In New Electron Correlation Methods and Their Applications, and Use of Atomic Orbitals with Exponential Asymptotes; Elsevier: Amsterdam, The Netherlands, 2021; pp. 107–123. [Google Scholar] [CrossRef]
  44. Smits, O.R.; Mewes, J.M.; Jerabek, P.; Schwerdtfeger, P. Oganesson: A Noble Gas Element That Is Neither Noble Nor a Gas. Angew. Chem. Int. Ed. 2020, 59, 23636–23640. [Google Scholar] [CrossRef]
  45. Sato, T.; Asai, M.; Borschevsky, A.; Stora, T.; Sato, N.; Kaneya, Y.; Tsukada, K.; Düllmann, C.E.; Eberhardt, K.; Eliav, E.; et al. Measurement of the first ionization potential of lawrencium, element 103. Nature 2015, 520, 209–211. [Google Scholar] [CrossRef] [Green Version]
  46. Sato, T.K.; Asai, M.; Borschevsky, A.; Beerwerth, R.; Kaneya, Y.; 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. Am. Chem. Soc. 2018, 140, 14609–14613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Kozlov, M.; Safronova, M.; López-Urrutia, J.C.; Schmidt, P. Highly charged ions: Optical clocks and applications in fundamental physics. Rev. Mod. Phys. 2018, 90, 045005. [Google Scholar] [CrossRef] [Green Version]
  48. Porsev, S.G.; Safronova, U.I.; Safronova, M.S.; Schmidt, P.O.; Bondarev, A.I.; Kozlov, M.G.; Tupitsyn, I.I.; Cheung, C. Optical clocks based on the Cf15+ and Cf17+ ions. Phys. Rev. A 2020, 102, 012802. [Google Scholar] [CrossRef]
  49. Kramida, A.; Ralchenko, Y.; Reader, J.; NIST ASD Team. NIST Atomic Spectra Database (Ver. 5.9); National Institute of Standards and Technology: Gaithersburg, MD, USA, 2021. Available online: https://physics.nist.gov/asd (accessed on 26 April 2022).
  50. Kramida, A. Update of Atomic Data for the First Three Spectra of Actinium. Atoms 2022, 10, 42. [Google Scholar] [CrossRef]
  51. Block, M.; Laatiaoui, M.; Raeder, S. Recent progress in laser spectroscopy of the actinides. Prog. Part. Nucl. Phys. 2021, 116, 103834. [Google Scholar] [CrossRef]
  52. Chhetri, P.; Ackermann, D.; Backe, H.; Block, M.; Cheal, B.; Droese, C.; Düllmann, C.E.; Even, J.; Ferrer, R.; Giacoppo, F.; et al. Precision Measurement of the First Ionization Potential of Nobelium. Phys. Rev. Lett. 2018, 120, 263003. [Google Scholar] [CrossRef]
  53. Warbinek, J.; Andelic, B.; Block, M.; Chhetri, P.; Claessens, A.; Ferrer, R.; Giacoppo, F.; Kaleja, O.; Kieck, T.; Kim, E.; et al. Advancing Radiation-Detected Resonance Ionization towards Heavier Elements and More Exotic Nuclides. Atoms 2022, 10, 41. [Google Scholar] [CrossRef]
  54. Laatiaoui, M.; Buchachenko, A.A.; Viehland, L.A. Laser Resonance Chromatography of Superheavy Elements. Phys. Rev. Lett. 2020, 125, 023002. [Google Scholar] [CrossRef]
  55. Laatiaoui, M.; Buchachenko, A.A.; Viehland, L.A. Exploiting transport properties for the detection of optical pumping in heavy ions. Phys. Rev. A 2020, 102, 013106. [Google Scholar] [CrossRef]
  56. Romero-Romero, E.; Block, M.; Kim, E.; Nothhelfer, S.; Raeder, S.; Ramanantoanina, H.; Rickert, E.; Schneider, J.; Sikora, P.; Laatiaoui, M. A progress report on Laser Resonance Chromatography. Atoms 2022. manuscript in preparation. [Google Scholar]
  57. Rickert, E.; Backe, H.; Block, M.; Laatiaoui, M.; Lauth, W.; Raeder, S.; Schneider, J.; Schneider, F. Ion Mobilities for Heaviest Element Identification. Hyperfine Interact. 2020, 241, 49. [Google Scholar] [CrossRef] [Green Version]
  58. Cheal, B.; Flanagan, K.T. Progress in laser spectroscopy at radioactive ion beam facilities. J. Phys. G Nucl. Part. Phys. 2010, 37, 113101. [Google Scholar] [CrossRef]
  59. Campbell, P.; Moore, I.; Pearson, M. Laser spectroscopy for nuclear structure physics. Prog. Part. Nucl. Phys. 2016, 86, 127–180. [Google Scholar] [CrossRef]
  60. Voss, A.; Sonnenschein, V.; Campbell, P.; Cheal, B.; Kron, T.; Moore, I.D.; Pohjalainen, I.; Raeder, S.; Trautmann, N.; Wendt, K. High-resolution laser spectroscopy of long-lived plutonium isotopes. Phys. Rev. A 2017, 95, 032506. [Google Scholar] [CrossRef] [Green Version]
  61. Raggio, A.; Pohjalainen, I.; Moore, I.D. Observation of Collisional De-Excitation Phenomena in Plutonium. Atoms 2022, 10, 40. [Google Scholar] [CrossRef]
  62. Zhang, K.; Studer, D.; Weber, F.; Gadelshin, V.M.; Kneip, N.; Raeder, S.; Budker, D.; Wendt, K.; Kieck, T.; Porsev, S.G.; et al. Detection of the Lowest-Lying Odd-Parity Atomic Levels in Actinium. Phys. Rev. Lett. 2020, 125, 073001. [Google Scholar] [CrossRef]
  63. Verstraelen, E.; Teigelhöfer, A.; Ryssens, W.; Ames, F.; Barzakh, A.; Bender, M.; Ferrer, R.; Goriely, S.; Heenen, P.H.; Huyse, M.; et al. Search for octupole-deformed actinium isotopes using resonance ionization spectroscopy. Phys. Rev. C 2019, 100, 044321. [Google Scholar] [CrossRef] [Green Version]
  64. 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. Nat. Commun. 2017, 8, 14520. [Google Scholar] [CrossRef] [Green Version]
  65. Granados, C.; Creemers, P.; Ferrer, R.; Gaffney, L.P.; Gins, W.; de Groote, R.; Huyse, M.; Kudryavtsev, Y.; Martínez, Y.; Raeder, S.; et al. In-gas laser ionization and spectroscopy of actinium isotopes near the N = 126 closed shell. Phys. Rev. C 2017, 96, 054331. [Google Scholar] [CrossRef] [Green Version]
  66. Weber, F.; Düllmann, C.E.; Gadelshin, V.; Kneip, N.; Oberstedt, S.; Raeder, S.; Runke, J.; Mokry, C.; Thörle-Pospiech, P.; Studer, D.; et al. Probing the Atomic Structure of Californium by Resonance Ionization Spectroscopy. Atoms 2022, 10, 51. [Google Scholar] [CrossRef]
  67. Nothhelfer, S.; Albrecht-Schönzart, T.E.; Block, M.; Chhetri, P.; Düllmann, C.E.; Ezold, J.G.; Gadelshin, V.; Gaiser, A.; Giacoppo, F.; Heinke, R.; et al. Nuclear structure investigations of Es253-255 by laser spectroscopy. Phys. Rev. C 2022, 105, l021302. [Google Scholar] [CrossRef]
  68. Raeder, S.; Ackermann, D.; Backe, H.; Beerwerth, R.; Berengut, J.; Block, M.; Borschevsky, A.; Cheal, B.; Chhetri, P.; Düllmann, C.E.; et al. Probing Sizes and Shapes of Nobelium Isotopes by Laser Spectroscopy. Phys. Rev. Lett. 2018, 120, 232503. [Google Scholar] [CrossRef] [Green Version]
  69. Kudryavtsev, Y.; Ferrer, R.; Huyse, M.; den Bergh, P.V.; Duppen, P.V. The in-gas-jet laser ion source: Resonance ionization spectroscopy of radioactive atoms in supersonic gas jets. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2013, 297, 7–22. [Google Scholar] [CrossRef] [Green Version]
  70. Zadvornaya, A.; Creemers, P.; Dockx, K.; Ferrer, R.; Gaffney, L.; Gins, W.; Granados, C.; Huyse, M.; Kudryavtsev, Y.; Laatiaoui, M.; et al. Characterization of Supersonic Gas Jets for High-Resolution Laser Ionization Spectroscopy of Heavy Elements. Phys. Rev. X 2018, 8, 041008. [Google Scholar] [CrossRef] [Green Version]
  71. Ferrer, R.; Verlinde, M.; Verstraelen, E.; Claessens, A.; Huyse, M.; Kraemer, S.; Kudryavtsev, Y.; Romans, J.; den Bergh, P.V.; Duppen, P.V.; et al. Hypersonic nozzle for laser-spectroscopy studies at 17 K characterized by resonance-ionization-spectroscopy-based flow mapping. Phys. Rev. Res. 2021, 3, 043041. [Google Scholar] [CrossRef]
  72. Piot, J. Studying Nuclear Structure at the extremes with S3. EPJ Web Conf. 2018, 178, 02027. [Google Scholar] [CrossRef] [Green Version]
  73. Ferrer, R.; Bastin, B.; Boilley, D.; Creemers, P.; Delahaye, P.; Liénard, E.; Fléchard, X.; Franchoo, S.; Ghys, L.; Huyse, M.; et al. In gas laser ionization and spectroscopy experiments at the Superconducting Separator Spectrometer (S3): Conceptual studies and preliminary design. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2013, 317, 570–581. [Google Scholar] [CrossRef] [Green Version]
  74. Romans, J.; Ajayakumar, A.; Authier, M.; Boumard, F.; Caceres, L.; Cam, J.F.; Claessens, A.; Damoy, S.; Delahaye, P.; Desrues, P.; et al. First Offline Results from the S3 Low-Energy Branch. Atoms 2022, 10, 21. [Google Scholar] [CrossRef]
  75. Papadakis, P.; Liimatainen, J.; Sarén, J.; Moore, I.; Eronen, T.; Partanen, J.; Pohjalainen, I.; Rinta-Antila, S.; Tuunanen, J.; Uusitalo, J. The MARA-LEB ion transport system. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2020, 463, 286–289. [Google Scholar] [CrossRef]
  76. Raeder, S.; Block, M.; Chhetri, P.; Ferrer, R.; Kraemer, S.; Kron, T.; Laatiaoui, M.; Nothhelfer, S.; Schneider, F.; Duppen, P.V.; et al. A gas-jet apparatus for high-resolution laser spectroscopy on the heaviest elements at SHIP. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2020, 463, 272–276. [Google Scholar] [CrossRef]
  77. Münzberg, D.; Block, M.; Claessens, A.; Ferrer, R.; Laatiaoui, M.; Lantis, J.; Nothhelfer, S.; Raeder, S.; Van Duppen, P. Resolution characterizations of JetRIS in Mainz using 164Dy. Atoms 2022, 10, 57. [Google Scholar] [CrossRef]
  78. Verlinde, M.; Ferrer, R.; Claessens, A.; Granados, C.A.; Kraemer, S.; Kudryavtsev, Y.; Li, D.; den Bergh, P.V.; Duppen, P.V.; Verstraelen, E. Single-longitudinal-mode pumped pulsed-dye amplifier for high-resolution laser spectroscopy. Rev. Sci. Instrum. 2020, 91, 103002. [Google Scholar] [CrossRef] [PubMed]
  79. Raeder, S.; Ferrer, R.; Granados, C.; Huyse, M.; Kron, T.; Kudryavtsev, Y.; Lecesne, N.; Piot, J.; Romans, J.; Savajols, H.; et al. Performance of Dye and Ti:sapphire laser systems for laser ionization and spectroscopy studies at S3. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2020, 463, 86–95. [Google Scholar] [CrossRef]
  80. Sonnenschein, V.; Ohashi, M.; Tomita, H.; Iguchi, T. A direct diode pumped continuous-wave Ti:sapphire laser as seed of a pulsed amplifier for high-resolution resonance ionization spectroscopy. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2020, 463, 512–514. [Google Scholar] [CrossRef]
  81. Dobaczewski, J.; Engel, J.; Kortelainen, M.; Becker, P. Correlating Schiff Moments in the Light Actinides with Octupole Moments. Phys. Rev. Lett. 2018, 121, 232501. [Google Scholar] [CrossRef] [Green Version]
  82. Ruiz, R.F.G.; Berger, R.; Billowes, J.; Binnersley, C.L.; Bissell, M.L.; Breier, A.A.; Brinson, A.J.; Chrysalidis, K.; Cocolios, T.E.; Cooper, B.S.; et al. Spectroscopy of short-lived radioactive molecules. Nature 2020, 581, 396–400. [Google Scholar] [CrossRef]
  83. von der Wense, L.; Seiferle, B.; Laatiaoui, M.; Neumayr, J.B.; Maier, H.J.; Wirth, H.F.; Mokry, C.; Runke, J.; Eberhardt, K.; Düllmann, C.E.; et al. Direct detection of the 229Th nuclear clock transition. Nature 2016, 533, 47–51. [Google Scholar] [CrossRef] [Green Version]
  84. Seiferle, B.; von der Wense, L.; Bilous, P.V.; Amersdorffer, I.; Lemell, C.; Libisch, F.; Stellmer, S.; Schumm, T.; Düllmann, C.E.; Pálffy, A.; et al. Energy of the 229Th nuclear clock transition. Nature 2019, 573, 243–246. [Google Scholar] [CrossRef] [Green Version]
  85. Sikorsky, T.; Geist, J.; Hengstler, D.; Kempf, S.; Gastaldo, L.; Enss, C.; Mokry, C.; Runke, J.; Düllmann, C.E.; Wobrauschek, P.; et al. Measurement of the Th229 Isomer Energy with a Magnetic Microcalorimeter. Phys. Rev. Lett. 2020, 125, 142503. [Google Scholar] [CrossRef]
  86. Peik, E.; Schumm, T.; Safronova, M.S.; Pálffy, A.; Weitenberg, J.; Thirolf, P.G. Nuclear clocks for testing fundamental physics. Quantum Sci. Technol. 2021, 6, 034002. [Google Scholar] [CrossRef]
  87. Thirolf, P.G.; Seiferle, B.; Wense, L.V. The Thorium-Isomer: Heartbeat for a Nuclear Clock. Nucl. Phys. News 2021, 31, 13–18. [Google Scholar] [CrossRef]
  88. Beeks, K.; Sikorsky, T.; Schumm, T.; Thielking, J.; Okhapkin, M.V.; Peik, E. The thorium-229 low-energy isomer and the nuclear clock. Nat. Rev. Phys. 2021, 3, 238–248. [Google Scholar] [CrossRef]
  89. Verlinde, M.; Kraemer, S.; Moens, J.; Chrysalidis, K.; Correia, J.G.; Cottenier, S.; Witte, H.D.; Fedorov, D.V.; Fedosseev, V.N.; Ferrer, R.; et al. Alternative approach to populate and study the Th229 nuclear clock isomer. Phys. Rev. C 2019, 100, 024315. [Google Scholar] [CrossRef] [Green Version]
  90. Seiferle, B.; Moritz, D.; Scharl, K.; Ding, S.; Zacherl, F.; Löbell, L.; Thirolf, P.G. Extending Our Knowledge about the 229Th Nuclear Isomer. Atoms 2022, 10, 24. [Google Scholar] [CrossRef]
  91. Seiferle, B.; von der Wense, L.; Thirolf, P.G. Lifetime Measurement of the Th229 nuclear isomer. Phys. Rev. Lett. 2017, 118, 042501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Tkalya, E.V.; Si, R. Internal conversion of the low-energy Th229m isomer in the thorium anion. Phys. Rev. C 2020, 101, 054602. [Google Scholar] [CrossRef]
  93. Campbell, C.J.; Radnaev, A.G.; Kuzmich, A. Wigner Crystals of 229Th for Optical Excitation of the Nuclear Isomer. Phys. Rev. Lett. 2011, 106, 223001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Thielking, J.; Okhapkin, M.V.; Głowacki, P.; Meier, D.M.; von der Wense, L.; Seiferle, B.; Düllmann, C.E.; Thirolf, P.G.; Peik, E. Laser spectroscopic characterization of the nuclear-clock isomer 229mTh. Nature 2018, 556, 321–325. [Google Scholar] [CrossRef]
  95. Hotchkis, M.; Child, D.; Zorko, B. Actinides AMS for nuclear safeguards and related applications. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2010, 268, 1257–1260. [Google Scholar] [CrossRef]
  96. Bosco, H.; Hamann, L.; Kneip, N.; Raiwa, M.; Weiss, M.; Wendt, K.; Walther, C. New horizons in microparticle forensics: Actinide imaging and detection of 238 Pu and 242m Am in hot particles. Sci. Adv. 2021, 7, 44. [Google Scholar] [CrossRef]
  97. Kneip, N.; Düllmann, C.E.; Gadelshin, V.; Heinke, R.; Mokry, C.; Raeder, S.; Runke, J.; Studer, D.; Trautmann, N.; Weber, F.; et al. Highly selective two-step laser ionization schemes for the analysis of actinide mixtures. Hyperfine Interact. 2020, 241, 45. [Google Scholar] [CrossRef] [Green Version]
  98. Raeder, S.; Kneip, N.; Reich, T.; Studer, D.; Trautmann, N.; Wendt, K. Recent developments in resonance ionization mass spectrometry for ultra-trace analysis of actinide elements. Radiochim. Acta 2019, 107, 645–652. [Google Scholar] [CrossRef]
  99. Liu, Y.; Stracener, D. High efficiency resonance ionization of thorium. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2020, 462, 95–101. [Google Scholar] [CrossRef]
  100. Galindo-Uribarri, A.; Liu, Y.; Romero, E.R.; Stracener, D.W. High efficiency laser resonance ionization of plutonium. Sci. Rep. 2021, 11, 23432. [Google Scholar] [CrossRef]
  101. Schönenbach, D.; Berg, F.; Breckheimer, M.; Hagenlocher, D.; Schönberg, P.; Haas, R.; Amayri, S.; Reich, T. Development, characterization, and first application of a resonant laser secondary neutral mass spectrometry setup for the research of plutonium in the context of long-term nuclear waste storage. Anal. Bioanal. Chem. 2021, 413, 3987–3997. [Google Scholar] [CrossRef]
  102. Savina, M.R.; Isselhardt, B.H.; Trappitsch, R. Simultaneous Isotopic Analysis of U, Pu, and Am in Spent Nuclear Fuel by Resonance Ionization Mass Spectrometry. Anal. Chem. 2021, 93, 9505–9512. [Google Scholar] [CrossRef]
  103. Dockx, K.; Cocolios, T.E.; Stora, T. ISOL Technique for the Production of 225Ac at CERN-MEDICIS. J. Med Imaging Radiat. Sci. 2019, 50, S92. [Google Scholar] [CrossRef]
  104. Duchemin, C.; Ramos, J.P.; Stora, T.; Ahmed, E.; Aubert, E.; Audouin, N.; Barbero, E.; Barozier, V.; Bernardes, A.P.; Bertreix, P.; et al. CERN-MEDICIS: A Review Since Commissioning in 2017. Front. Med. 2021, 8. [Google Scholar] [CrossRef]
Figure 1. Logo of the virtual workshop on the Atomic Structure of Actinides & Related Topics, 26–28 May 2021.
Figure 1. Logo of the virtual workshop on the Atomic Structure of Actinides & Related Topics, 26–28 May 2021.
Atoms 10 00061 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Laatiaoui, M.; Raeder, S. New Developments in the Production and Research of Actinide Elements. Atoms 2022, 10, 61. https://doi.org/10.3390/atoms10020061

AMA Style

Laatiaoui M, Raeder S. New Developments in the Production and Research of Actinide Elements. Atoms. 2022; 10(2):61. https://doi.org/10.3390/atoms10020061

Chicago/Turabian Style

Laatiaoui, Mustapha, and Sebastian Raeder. 2022. "New Developments in the Production and Research of Actinide Elements" Atoms 10, no. 2: 61. https://doi.org/10.3390/atoms10020061

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop