Next Article in Journal
From Sub-Solar to Super-Solar Chemical Abundances along the Quasar Main Sequence
Previous Article in Journal
Surface Scattering Expansion of the Casimir–Polder Interaction for Magneto-Dielectric Bodies: Convergence Properties for Insulators, Conductors, and Semiconductors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Towards Precision Muonic X-ray Measurements of Charge Radii of Light Nuclei

by
Ben Ohayon
1,*,
Andreas Abeln
2,
Silvia Bara
3,
Thomas Elias Cocolios
3,
Ofir Eizenberg
1,
Andreas Fleischmann
2,
Loredana Gastaldo
2,
César Godinho
4,5,
Michael Heines
3,
Daniel Hengstler
2,
Guillaume Hupin
5,
Paul Indelicato
6,
Klaus Kirch
7,8,
Andreas Knecht
8,
Daniel Kreuzberger
2,
Jorge Machado
4,
Petr Navratil
9,
Nancy Paul
6,*,
Randolf Pohl
10,11,
Daniel Unger
2,
Stergiani Marina Vogiatzi
8,
Katharina von Schoeler
7,8 and
Frederik Wauters
10,12
add Show full author list remove Hide full author list
1
Physics Department, Technion—Israel Institute of Technology, Haifa 3200003, Israel
2
Kirchhoff Institut für Physik, Universität Heidelberg, Im Neuenheimer Feld 227, 69120 Heidelberg, Germany
3
Katholieke Universiteit (KU) Leuven, Instituut voor Kern- en Stralingsfysica, 3001 Leuven, Belgium
4
Departamento de Física da Faculdade de Ciências e Tecnologia, Universidade Nova de Lisboa, Monte da Caparica, 2892-516 Caparica, Portugal
5
Le Laboratoire de Physique des Deux Infinis Irène Joliot-Curie (IJCLab), Centre National de la Recherche Scientifique/Institut National de Physique Nucléaire et de Physique des Particules (CNRS/IN2P3), Université Paris-Saclay, 91405 Orsay, France
6
Laboratoire Kastler Brossel, Sorbonne Université, CNRS, L’École-Normale Supérieure, Paris Sciences et Lettres (ENS-PSL) Research University, Collège de France, Case 74, 4, Place Jussieu, 75005 Paris, France
7
Institute for Particle Physics and Astrophysics, Eidgenössische Technische Hochschule (ETH) Zürich, 8093 Zürich, Switzerland
8
Paul Scherrer Institute, 5232 Villigen, Switzerland
9
TRIUMF, 4004 Wesbrook Mall, Vancouver, BC V6T 2A3, Canada
10
PRISMA+ Cluster of Excellence, Johannes Gutenberg-Universität Mainz, 55128 Mainz, Germany
11
Institut für Physik, QUANTUM, Johannes Gutenberg-Universität Mainz, 55128 Mainz, Germany
12
Institut für Kernphysik, Johannes Gutenberg-Universität Mainz, 55128 Mainz, Germany
*
Authors to whom correspondence should be addressed.
Physics 2024, 6(1), 206-215; https://doi.org/10.3390/physics6010015
Submission received: 1 December 2023 / Revised: 8 January 2024 / Accepted: 15 January 2024 / Published: 17 February 2024
(This article belongs to the Special Issue Precision Physics and Fundamental Physical Constants (FFK 2023))

Abstract

:
We, the QUARTET Collaboration, propose an experiment to measure the nuclear charge radii of light elements with up to 20 times higher accuracy. These are essential both for understanding nuclear physics at low energies, and for experimental and theoretical applications in simple atomic systems. Such comparisons advance the understanding of bound-state quantum electrodynamics and are useful for searching for new physics beyond the Standard Model. The energy levels of muonic atoms are highly susceptible to nuclear structure, especially to the mean square charge radius. The radii of the lightest nuclei (with the atomic number, Z = 1 , 2 ) have been determined with high accuracy using laser spectroscopy in muonic atoms, while those of medium mass and above were determined using X-ray spectroscopy with semiconductor detectors. In this communication, we present a new experiment, aiming to obtain precision measurements of the radii of light nuclei 3 Z 10 using single-photon energy measurements with cryogenic microcalorimeters; a quantum-sensing technology capable of high efficiency with outstanding resolution for low-energy X-rays.

1. Introduction

Muonic atoms are highly suitable systems for studying the nucleus. Due to the heavy mass of muons ( m μ 200 m e , with m e the electron mass), the Bohr radius of muonic atoms is approximately 200 times smaller than that of electronic atoms, and thus, for low angular momentum states, the muon wavefunction has a 200 3   10 6 times larger overlap with that of the nucleus. The nuclear properties thus lead to measurable shifts in the atomic transition energies, making muonic atom spectroscopy an effective probe of phenomena such as finite nuclear size effects [1,2,3,4,5,6,7], relativistic QED (quantum electrodynamics) contributions [4,7,8,9], and possible short-range interactions carried by new mediators [10,11,12,13,14,15,16]. These systems are particularly well-suited to accurately determining the RMS (root mean squre) nuclear charge radius (the slope of the Sachs form factor at small momentum transfer, henceforth called ‘radius’ for brevity), which can be obtained using the spectroscopy of low-lying radiative transitions (mostly 2 P 1 S ) [1,17,18]. Indeed, historically, the best measurements of absolute radii have been obtained using muonic atom spectroscopy, sometimes leading to unexpected results such as the ‘proton radius puzzle’ [19,20,21,22,23,24,25].
The radius is a fundamental property of the nucleus, and knowledge of it is not only significant in the development of a nuclear structure theory, but also for obtaining a reliable comparison between the experimental results and theoretical expectations at the accuracy frontier of Standard Model tests with atoms and nuclei [3]. Accordingly, the radius of the proton and deuteron are considered fundamental constants, on a similar footing as their masses [26], with those of heavier nuclei expected to be included in the next CODATA adjustment of fundamental constants.
In Table 1, we collect the most precise values for the radii of light and stable nuclei. It is immediately seen that the elements in the most critical need of improved radii are those with a nuclear charge, Z, just above helium, a region that is currently beyond the reach of laser spectroscopy, where solid-state detectors are the most unsuitable. Advantageously, these nuclei are also those whose structure can now be calculated by the most advanced ab initio nuclear theory methods, as detailed in what follows.

2. Physics Cases

2.1. Nuclear Structure

In contemporary ab initio approaches, nucleon–nucleon and three-nucleon interactions are derived from chiral effective field theory (EFT) and used to calculate observables in quantum many-body systems with quantifiable uncertainties [43,44]. Due to a combinatorial increase in computational cost with mass number, A, high-precision calculations are limited to A 16 . The next-generation calculations also treats the coupling to the continuum and what is referred to as “open quantum systems” [45,46,47], which are crucial for accurately reproducing the structure of both light and exotic nuclei, especially their spatial extension.
While nuclear forces are fit for the properties of light, dilute nuclear systems, testing their ability to predict the properties of A > 4 bound systems (i.e., when nuclear density achieves saturation but A remains sufficiently low for highly precise calculations) provides a means to gauge the measurements accuracy. The test allows for an investigation of the precision of the chiral EFT itself, ultimately challenging the understanding of the strong force at low energies. Nuclear radii are a particularly interesting testing ground, as they can be calculated to a high level of precision when the coupling with continuum is included, and the measurements of absolute charge radii may even be used to obtain, e.g., electric quadrupole (E2) observables by applying the observed correlations [48].
As an example, we consider the radii of the lithium isotopic chain, whose isotopic differences relative to 6Li were extracted with high precision from optical isotope shift measurements [36,49,50]. Different ab initio nuclear models have been put forward to reproduce the results. However, the results cannot distinguish, because of the dominating uncertainty in the 6Li radius, to which the chain is referenced. The same happens in the beryllium [51] and boron [39] cases. The corresponding measurement goal is to distinguish well which model reproduces the measured radii within a few 10 3 , of its value, which can be reached at the early stage of an experiment.
A more demanding accuracy is required when considering differential observables, such as the differences between the radii of mirror nuclei (the nuclei with neutron and proton numbers being interchanged): Δ r mir , which are a focus of contemporary studies of nuclear structure (see e.g., [52,53,54,55,56,57]). A linear relationship between Δ r mir and neutron skins [58,59,60], which are particularly complicated to directly measure, has been found. Measuring Δ r mir can thus contribute to the understanding of the variations in neutron skin with isospin, while any deviations from a linearity would indicate the role that continuum degrees of freedom play in exotic nuclei structure evolution. Light mirror pairs such as 7Li-7Be and 8Li-8B possess a large isospin asymmetry and are, hence, well-suited to testing these theoretical predictions. Differences in radii were measured with the optical isotope shifts in the Li [36,49,50] and Be [51] chains, while the measurement of 8B [61] is ongoing. A need to measure Δ r mir accurately enough to not be limited by the reference radii sets a stringent demand to measure the radii with the accuracy as high as 10 3 for a stable isotope of each of the three elements.

2.2. QED and Beyond Standard Model (BSM)

The cases discussed in Section 2.1, are rooted in nuclear physics. Here, we show that high-precision measurements of the radii are necessary for testing the QED effects at the frontier of research in atomic physics. There are two main approaches to performing precision atomic structure calculations. The first is based on a perturbative expansion with respect to relativistic and QED effects in the Coulomb field, with electron–electron correlations being treated non-perturbatively. This approach is best adapted for low-Z systems such as hydrogen and helium. The second approach treats the QED and relativistic effects of all orders in the Coulomb field. It is needed for high-Z atoms, but loses accuracy in the low-Z region, where the correlations are more pronounced. The intermediate region, Z 6 is particularly interesting, as both approaches have each maximum uncertainty. There is therefore an interest in studying QED for that light, but not too light, few-electron systems (see, e.g., [62,63,64] and the discussion therein).
Precise nuclear radii are needed to disentangle potentially missing QED contributions as a function of Z. To illustrate this point, let us consider 2 3 S 1 2 3 P j transitions in 7Li+. Their hyperfine-averaged value was determined to within 0.4 MHz [65]. From this measurement, the two-electron Lamb shift could be experimentally determined with 1.5 MHz precision, dominated by the uncertainty in the nuclear radius. Considering the ongoing experiments aim to obtain an accuracy of the order of 100 kHz [66], an order-of-magnitude improvement in the radius would allow for the missing α 8 ( α denotes the fine structure constant) theory contribution of the order 3 MHz, to be determined with an uncertainty of 10 % [64]. Moreover, this effect can be downscaled to the analogous transition in the helium atom and to shed light on the observed, and recently confirmed, deviations between experiment and theory [67,68].
Above lithium, a new experimental program at Technical University Darmstadt, Germany, focuses on transitions in helium-like ions (HLIs) from beryllium to nitrogen [69,70,71,72], with the measurements in 12C already completed [73,74]. Improved radii are crucial for confronting the results of the Darmstadt campaign with the state-of-the-art calculations, especially if one considers using measurements at a high Z to determine the missing QED contributions, and using these calculations in the measurements with lower Z.
At the accuracy frontier of BSM physics searches, new interactions are hunted through their manifestation as significant differences between experiment and theory. Currently, the strongest bound on fifth forcesbetween charged leptons and neutrons derives from a combination of muonic and electronic isotope shifts in the hydrogen–deuterium pair [13]. When the mediating bosons are heavy, the sensitivity scale is Z 3 , favoring highly charged systems [75]. To utilize high-precision measurements of optical isotope shifts in simple enough electronic systems for BSM tests, one needs to considerably improve muonic isotope shifts. Accordingly, and as both the nuclear theory and experimental uncertainties associated with calibration largely cancel in the difference, we consider to measure these isotope shifts with suitable precision to determine differential radii with an accuracy above 10 3 fm, limited by the residual nuclear theory uncertainty.
We conclude this Section by noting that improved measurements of transitions to the ground level in a muonic atom directly translate to a better prediction of the muonic atom Lamb shift, which is accessible to laser spectroscopy. This statement is true irrespective of the nuclear-structure uncertainty. Quantitatively, a few parts per million measurement of the 2 P 1 S energy in muonic lithium translates to a few-meV prediction of the muonic lithium Lamb shift. Such a narrow search region greatly reduces the time needed to conduct a successful high-precision laser spectroscopy measurement, thus increasing the feasibility of the experiments suggested in Ref. [76]. The resolution afforded by laser spectroscopy in muonic atoms would, in turn, enable the hyperfine structure to be resolved and determine the Zemach radius (a convolution of the electric and magnetic distributions) of lithium isotopes. This determination is highly demanded due to the sharp disagreement between this value, as calculated by nuclear theory and as determined by the electronic measurements [77,78,79,80]. Moreover, ongoing work suggests that the redundancy between X-ray and laser measurements in the same muonic species constitutes a powerful platform to search for new physics carried by new medium-mass (of the order of MeV) bosons.

3. Theory Considerations

The energy of the atomic transition between principle quantum numbers can be written as
E = E D + δ E QED + δ E FNS + δ E TPE + ,
where E D is the Dirac energy for a point-like nucleus, δ E QED is the sum of leading quantum electrodynamics effects, δ E FNS r c 2 is the leading order correction due to the finite nuclear size, from which the charge radius is extracted, and δ E TPE is the sum of corrections stemming from the two-photon exchange, which depends on the nuclear structure, namely the nuclear polarizability and higher charge moments of the nucleus. At the precision level that is foreseen for the project considered here, the uncertainty in point-nucleus QED corrections is negligible [81,82]. Accordingly, once the experimental accuracy of the transition energies is improved, the uncertainty in δ E TPE and higher-order nuclear structure contributions [29] is expected to dominate the derived radii. Based on the calculations for the lightest nuclei (see [28] and references therein), preliminary results for 6,7Li [83] and the recent studies on heavier systems [84], we estimate that a 5 to 10 % uncertainty in the calculated δ E TPE is achievable, resulting in an absolute radii with an accuracy of the order of a few times 10 4 , similar to that for the neighboring nuclei (see Table 1). Further calculations for the nuclei of interest are in progress. These could be achieved by applying the no-core shell model [85] with the Lanczos method [86,87]. In addition to the calculations, the helpful information on the nuclear shape can, in some cases (notably 12C [88]), be incorporated from elastic electron scattering measurements.
More accurate atomic theory considerations are needed in order to account for the unresolved fine and hyperfine structure features, mixed finite size and QED corrections, and shifts from spectator electrons, which screen the nuclear potential seen by the muon [89,90]. Accordingly, we calculated the atomic structure of the targeted systems using the Multiconfiguration Dirac-Fock General Matrix Elements (MCDFGME) code, which is able evaluate the energies, transition probabilities, and hyperfine structure for exotic atoms composed of a nucleus, an arbitrary number of electrons, and an additional fermion or boson [91,92,93,94,95]. The energies are obtained using a full-atomic wave function composed of a determinant with all the electrons, multiplied by the muon wave function, and by solving the full coupled system of differential equations. The electron–electron and muon–electron interactions are chosen to represent the full Breit operator with Coulomb, magnetic, and retardation in the Coulomb gauge. Nuclear deformation effects could also be of importance [96,97,98,99,100]. They are not explicitly included in the atomic structure calculation and will be evaluated separately for this work.
Based on successful studies in the lightest systems, we are confident that precision experiments with these heavier muonic atoms will instigate new activity in the relevant atomic and nuclear theories, potentially contributing to related fields, such as studies with antiprotonic atoms [101] and highly charged ions [102].

4. Experimental Considerations

The radii of most of the stable nuclei were measured using traditional muonic X-ray spectroscopy with semiconductor detectors [103]. However, due to their moderate resolving power (fractional resolution) below 200 keV, compounded with the Z 2 scaling of the fractional contribution of the radius to the energy levels, this approach is insufficient to precisely determine finite-size effects in light nuclei. In contrast with the semiconducting detectors, crystal spectrometers offer high resolution in the multi-keV regime [104]. This detection method was used to determine the 2 P 1 S transition energy in μ 12 C to 5 ppm, and derive the radius with an accuracy of 2 × 10 3 fm [41,105]. This demonstrates that an X-ray detector with a resolving power of a few thousand enables the radii measurements in light nuclei with a precision better than 10 3 . However, this method suffers from low efficiency and a narrow bandwidth, making it impractical to extend to the entire series of light muonic elements in the available facilities considering beamtime constraints.
In order to measure the relevant transitions in light elements with sufficient accuracy and within a reasonable time, we use a different technology: metallic magnetic calorimeters (MMCs) [106], operated at exceptionally low cryogenic temperatures (about 20 mK). This quantum-sensing single-photon energy-detection technique achieves a high resolving power of few thousand, well-understood nonlinearity, and high quantum efficiency, as follows: above 50% up to 40 keV, and above 5% up to 180 keV [107]. It is thus ideally suited to tackling the problem of measuring the charge radii of light elements ( Z 10 ) using X-ray spectroscopy in muonic atoms. The principle of detection is that an X-ray is absorbed in a metallic absorber and its complete energy is converted to a temperature increase. This leads to a magnetization change in a paramagnetic material connecting the absorber to a thermal bath. This change in magnetization is high-sensitively detected using a superconducting quantum interference device (SQUID).
The first proof-of-concept measurement is currently undergoing its preparation. We transported an existing micro-calorimeter array for X-ray spectroscopy (maXs) [108,109] in a sidearm of a dilution refrigerator, from the Kirchhoff Institute for Physics in Heidelberg, to a secondary muon beamline at the Paul Scherrer Institute (PSI), Villigen, Switzerland. The detector is planned to be integrated with the existing muon, electron, and photon detectors from the muX experiment [110], which is expected to allow identifying and suppressing different sources of background. For detailed information on the detector, its performance, and its integration with the beamline, see [107].
Precise absolute X-ray measurements not only require a high resolution, ample statistics, and effective background suppression, but also a robust calibration strategy. The calibration function of the MMC detectors is to be determined periodically by using readily available sources [111]. To complement this method, a commercial X-ray tube is designed to be placed in the experimental system. There, electrons excite various metallic targets, emitting their characteristic X-rays, some of which have energies that are known to sub-ppm accuracy [112,113], and traceable to the International System of Units’ second [114]. Similarly, the energy of muonic X-ray lines that do not involve an S-state can be reliably calculated to sub-ppm accuracy [115]. In this way, higher-lying lines in heavier muonic atoms could calibrate 2 P 1 S lines in lighter systems.
Finally, let us note that the experiment considered in this paper and other next-generation experiments on exotic atoms (e.g., [116,117]), could significantly benefit from the ppm-level absolute gamma-ray energy measurements in the 20–200 keV range, especially from readily available long-lived commercial sources.

5. Summary

QUARTET is a new Collaboration that seeks to significantly improve the experimental values of the charge radii from lithium to neon by means of a precision muonic X-ray spectroscopy with metallic magnetic calorimeters, filling the gap between laser spectroscopy- and semiconductor-based X-ray spectroscopy for the elements that are beyond reach of both the methods. It will be the first time that those detectors are used with exotic atoms. The expected ten-fold improvement in precision will significantly impact nuclear structure and QED tests, and pave the way for moving BSM physics searches through the combinations of the muonic and electronic isotope shifts on one side and the laser and X-ray spectroscopy, on the other side.

Author Contributions

Conceptualization, B.O., N.P., F.W., P.I., A.K., R.P. and K.K.; methodology, B.O., N.P., F.W., L.G., A.F., J.M., P.N., G.H., D.U., S.M.V., K.v.S. and K.K.; software, F.W., J.M., D.K., P.I., D.H., C.G., A.A., K.v.S. and M.H.; formal analysis, P.I., G.H. and P.N.; investigation, B.O., A.A., S.B., T.E.C., O.E., A.F., L.G., C.G., M.H., D.H., G.H., P.I., A.K., D.K., J.M., P.N., N.P., R.P., D.U., S.M.V., K.v.S. and F.W.; resources, B.O., N.P., T.E.C., A.F., L.G., P.I., K.K., J.M., P.N., R.P. and F.W.; data curation, F.W., B.O., O.E., D.U. and N.P.; writing—original draft preparation, B.O.; writing—review and editing, B.O., A.A., S.B., T.E.C., K.K., O.E., A.F., L.G., C.G., M.H., D.H., G.H., P.I., A.K., D.K., J.M., P.N., N.P., R.P., D.U., S.M.V., K.v.S. and F.W.; project administration, B.O. and N.P.; funding acquisition, B.O., N.P., F.W., R.P., T.E.C., L.G., A.F., A.K. and K.K.; All authors have read and agreed to the published version of the manuscript.

Funding

B.O. is thankful for the support of the Council for Higher Education Program for Hiring Outstanding Faculty Members in Quantum Science and Technology. The Kirchhoff Institute for Physics group at Heidelberg University is supported by Field Of Focus II initiative at Heidelberg University. D.U. acknowledges the support by the Research Training Group HighRR (GRK 2058) funded through the Deutsche Forschungsgemeinschaft, DFG. The work of the KU Leuven group is supported by FWO-Vlaanderen (Belgium), KU Leuven BOF C14/22/104, and European Research Council, grant no. 101088504 (NSHAPE). P.N. acknowledges support from the NSERC Grant No. SAPIN-2022-00019. TRIUMF receives federal funding via a contribution agreement with the National Research Council of Canada. The Lisboa group is supported in part by Fundação ra a Ciència e Tecnologia (FCT; Portugal) through research center Grant No. UID/FIS/04559/2020 to LIBPhys-UNL. The work of the ETH group was supported by the ETH Research Grant 22-2 ETH-023, Switzerland.

Data Availability Statement

All the data relevant to this communication is found in it.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wu, C.S.; Wilets, L. Muonic atoms and nuclear structure. Annu. Rev. Nucl. Sci. 1969, 19, 527–606. [Google Scholar] [CrossRef]
  2. Friar, J.L. Nuclear finite-size effects in light muonic atoms. Ann. Phys. 1979, 122, 151–196. [Google Scholar] [CrossRef]
  3. Friar, J.L. The structure of light nuclei and its effect on precise atomic measurements. In Precision Physics of Simple Atomic Systems; Karshenboim, S.G., Smirnov, V.B., Eds.; Springer: Berlin/Heidelberg, Germany, 2003; pp. 59–79. [Google Scholar] [CrossRef]
  4. Karshenboim, S.G. Precision physics of simple atoms: QED tests, nuclear structure and fundamental constants. Phys. Rep. 2005, 422, 1–63. [Google Scholar] [CrossRef]
  5. Kanda, S.; Ishida, K.; Iwasaki, M.; Ma, Y.; Okada, S.; Takamine, A.; Ueno, H.; Midorikawa, K.; Saito, N.; Wada, S.; et al. Measurement of the proton Zemach radius from the hyperfine splitting in muonic hydrogen atom. J. Phys. Conf. Ser. 2018, 1138, 012009. [Google Scholar] [CrossRef]
  6. Pizzolotto, C.; Adamczak, A.; Bakalov, D.; Baldazzi, G.; Baruzzo, M.; Benocci, R.; Bertoni, R.; Bonesini, M.; Bonvicini, V.; Cabrera, H.; et al. The FAMU experiment: Muonic hydrogen high precision spectroscopy studies. Eur. Phys. J. A 2020, 56, 185. [Google Scholar] [CrossRef]
  7. Antognini, A.; Bacca, S.; Fleischmann, A.; Gastaldo, L.; Hagelstein, F.; Indelicato, P.; Knecht, A.; Lensky, V.; Ohayon, B.; Pascalutsa, V.; et al. Muonic-atom spectroscopy and impact on nuclear structure and precision QED theory. arXiv 2022, arXiv:2210.16929. [Google Scholar] [CrossRef]
  8. Eides, M.I.; Grotch, H.; Shelyuto, V.A. Theory of Light Hydrogenic Bound States; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2007. [Google Scholar] [CrossRef]
  9. Antognini, A.; Lin, Y.H.; Meißner, U.G. Precision calculation of the recoil–finite-size correction for the hyperfine splitting in muonic and electronic hydrogen. Phys. Lett. B 2022, 835, 137575. [Google Scholar] [CrossRef]
  10. Karshenboim, S.G. Precision physics of simple atoms and constraints on a light boson with ultraweak coupling. Phys. Rev. Lett. 2010, 104, 220406. [Google Scholar] [CrossRef]
  11. Barger, V.; Chiang, C.W.; Keung, W.Y.; Marfatia, D. Proton size anomaly. Phys. Rev. Lett. 2011, 106, 153001. [Google Scholar] [CrossRef] [PubMed]
  12. Villalba-Chávez, S.; Golub, A.; Müller, C. Axion-modified photon propagator, Coulomb potential, and Lamb shift. Phys. Rev. D 2018, 98, 115008. [Google Scholar] [CrossRef]
  13. Delaunay, C.; Frugiuele, C.; Fuchs, E.; Soreq, Y. Probing new spin-independent interactions through precision spectroscopy in atoms with few electrons. Phys. Rev. D 2017, 96, 115002. [Google Scholar] [CrossRef]
  14. Jentschura, U.D.; Nándori, I. Atomic physics constraints on the X boson. Phys. Rev. A 2018, 97, 042502. [Google Scholar] [CrossRef]
  15. Karshenboim, S.G.; McKeen, D.; Pospelov, M. Constraints on muon-specific dark forces. Phys. Rev. D 2014, 90, 073004. [Google Scholar] [CrossRef]
  16. Claudia, F.; Clara, P. Muonic vs electronic dark forces: A complete EFT treatment for atomic spectroscopy. J. High Energy Phys. 2022, 2022, 002. [Google Scholar] [CrossRef]
  17. Engfer, R.; Schneuwly, H.; Vuilleumier, J.; Walter, H.; Zehnder, A. Charge-distribution parameters, isotope shifts, isomer shifts, and magnetic hyperfine constants from muonic atoms. At. Data Nucl. Data Tables 1974, 14, 509–597. [Google Scholar] [CrossRef]
  18. Fricke, G.; Bernhardt, C.; Heilig, K.; Schaller, L.; Schellenberg, L.; Shera, E.; Dejager, C. Nuclear ground state charge radii from electromagnetic interactions. At. Data Nucl. Data Tables 1995, 60, 177–285. [Google Scholar] [CrossRef]
  19. Lorenz, I.; Hammer, H.W.; Meißner, U.G. The size of the proton: Closing in on the radius puzzle. Eur. Phys. J. A 2012, 48, 151. [Google Scholar] [CrossRef]
  20. Pohl, R.; Gilman, R.; Miller, G.A.; Pachucki, K. Muonic hydrogen and the proton radius puzzle. Annu. Rev. Nucl. Part. Sci. 2013, 63, 175–204. [Google Scholar] [CrossRef]
  21. Carlson, C.E. The proton radius puzzle. Prog. Part. Nucl. Phys. 2015, 82, 59–77. [Google Scholar] [CrossRef]
  22. Hammer, H.W.; Meißner, U.G. The proton radius: From a puzzle to precision. Sci. Bull. 2020, 65, 257–258. [Google Scholar] [CrossRef] [PubMed]
  23. Peset, C.; Pineda, A.; Tomalak, O. The proton radius (puzzle?) and its relatives. Prog. Part. Nucl. Phys. 2021, 121, 103901. [Google Scholar] [CrossRef]
  24. Jentschura, U.D. Proton radius: A puzzle or a solution!? J. Phys. Conf. Ser. 2022, 2391, 012017. [Google Scholar] [CrossRef]
  25. Meißner, U.G. The proton radius and its relatives—much ado about nothing? J. Phys. Conf. Ser. 2023, 2586, 012006. [Google Scholar] [CrossRef]
  26. Tiesinga, E.; Mohr, P.J.; Newell, D.B.; Taylor, B.N. CODATA recommended values of the fundamental physical constants: 2018. Rev. Mod. Phys. 2021, 93, 025010. [Google Scholar] [CrossRef] [PubMed]
  27. Antognini, A.; Nez, F.; Schuhmann, K.; Amaro, F.D.; Biraben, F.; Cardoso, J.M.R.; Covita, D.S.; Dax, A.; Dhawan, S.; Diepold, M.; et al. Proton structure from the measurement of 2S-2P transition frequencies of muonic hydrogen. Science 2013, 339, 417–420. [Google Scholar] [CrossRef] [PubMed]
  28. Pachucki, K.; Lensky, V.; Hagelstein, F.; Li Muli, S.S.; Bacca, S.; Pohl, R. Comprehensive theory of the Lamb shift in light muonic atoms. arXiv 2022, arXiv:2212.13782. [Google Scholar] [CrossRef]
  29. Pachucki, K.; Patkós, V.; Yerokhin, V.A. Three-photon-exchange nuclear structure correction in hydrogenic systems. Phys. Rev. A 2018, 97, 062511. [Google Scholar] [CrossRef]
  30. Parthey, C.G.; Matveev, A.; Alnis, J.; Pohl, R.; Udem, T.; Jentschura, U.D.; Kolachevsky, N.; Hänsch, T.W. Precision measurement of the hydrogen-deuterium 1S − 2S isotope shift. Phys. Rev. Lett. 2010, 104, 233001. [Google Scholar] [CrossRef] [PubMed]
  31. Schuhmann, K.; et al. [The CREMA Collaboration] The helion charge radius from laser spectroscopy of muonic helium-3 ions. arXiv 2023, arXiv:2305.11679. [Google Scholar] [CrossRef]
  32. Krauth, J.J.; Schuhmann, K.; Ahmed, M.A.; Amaro, F.D.; Amaro, P.; Biraben, F.; Chen, T.L.; Covita, D.S.; Dax, A.J.; Diepold, M.; et al. Measuring the α-particle charge radius with muonic helium-4 ions. Nature 2021, 589, 527–531. [Google Scholar] [CrossRef] [PubMed]
  33. Suelzle, L.R.; Yearian, M.R.; Crannell, H. Elastic electron scattering from Li6 and Li7. Phys. Rev. 1967, 162, 992–1005. [Google Scholar] [CrossRef]
  34. Bumiller, F.A.; Buskirk, F.R.; Dyer, J.N.; Monson, W.A. Elastic electron scattering from 6Li and 7Li at low momentum transfer. Phys. Rev. C 1972, 5, 391–395. [Google Scholar] [CrossRef]
  35. Li, G.; Sick, I.; Whitney, R.; Yearian, M. High-energy electron scattering from 6Li. Nucl. Phys. A 1971, 162, 583–592. [Google Scholar] [CrossRef]
  36. Nörtershäuser, W.; Neff, T.; Sánchez, R.; Sick, I. Charge radii and ground state structure of lithium isotopes: Experiment and theory reexamined. Phys. Rev. C 2011, 84, 024307. [Google Scholar] [CrossRef]
  37. Nörtershäuser, W.; Sánchez, R.; Ewald, G.; Dax, A.; Behr, J.; Bricault, P.; Bushaw, B.A.; Dilling, J.; Dombsky, M.; Drake, G.W.F.; et al. Isotope-shift measurements of stable and short-lived lithium isotopes for nuclear-charge-radii determination. Phys. Rev. A 2011, 83, 012516. [Google Scholar] [CrossRef]
  38. Jansen, J.A.; Peerdeman, R.T.; De Vries, C. Nuclear charge radii of 12C and 9Be. Nucl. Phys. A 1972, 188, 337–352. [Google Scholar] [CrossRef]
  39. Maaß, B.; Hüther, T.; König, K.; Krämer, J.; Krause, J.; Lovato, A.; Müller, P.; Pachucki, K.; Puchalski, M.; Roth, R.; et al. Nuclear charge radii of 10,11B. Phys. Rev. Lett. 2019, 122, 182501. [Google Scholar] [CrossRef] [PubMed]
  40. Barnett, B.; Gyles, W.; Johnson, R.; Erdman, K.; Johnstone, J.; Kraushaar, J.; Lepp, S.; Masterson, T.; Rost, E.; Gill, D.; et al. Proton radii determinations from the ratio of π+ elastic scattering from 11B and 12C. Phys. Lett. B 1980, 97, 45–49. [Google Scholar] [CrossRef]
  41. Ruckstuhl, W.; Aas, B.; Beer, W.; Beltrami, I.; Bos, K.; Goudsmit, P.F.A.; Leisi, H.J.; Strassner, G.; Vacchi, A.; De Boer, F.W.N.; et al. Precision measurement of the 2p-1s transition in muonic 12C: Search for new muon-nucleon interactions or accurate determination of the rms nuclear charge radius. Nucl. Phys. A 1984, 430, 685–712. [Google Scholar] [CrossRef]
  42. de Boer, F.W.N.; Aas, B.; Baertschi, P.; Beer, W.; Beltrami, I.; Bos, K.; Goudsmit, P.F.A.; Kiebele, U.; Jeckelmann, B.; Leisi, H.J.; et al. Precision measurement of the 2p-1s transition wavelength in muonic 13C. Nucl. Phys. A 1985, 444, 589–596. [Google Scholar] [CrossRef]
  43. Epelbaum, E.; Hammer, H.-W.; Meißner, U.-G. Modern theory of nuclear forces. Rev. Mod. Phys. 2009, 81, 1773–1825. [Google Scholar] [CrossRef]
  44. Hergert, H. A Guided tour of ab initio nuclear many-body theory. Front. Phys. 2020, 8, 379. [Google Scholar] [CrossRef]
  45. Dohet-Eraly, J.; Navrátil, P.; Quaglioni, S.; Horiuchi, W.; Hupin, G.; Raimondi, F. 3He(α,γ)7Be and 3H(α,γ)7Li astrophysical S factors from the no-core shell model with continuum. Phys. Lett. B 2016, 757, 430–436. [Google Scholar] [CrossRef]
  46. Navrátil, P.; Quaglioni, S.; Hupin, G.; Romero-Redondo, C.; Calci, A. Unified ab initio approaches to nuclear structure and reactions. Phys. Scr. 2016, 91, 053002. [Google Scholar] [CrossRef]
  47. Vorabbi, M.; Navrátil, P.; Quaglioni, S.; Hupin, G. 7Be and 7Li nuclei within the no-core shell model with continuum. Phys. Rev. C 2019, 100, 024304. [Google Scholar] [CrossRef]
  48. Caprio, M.A.; Fasano, P.J.; Maris, P. Robust ab initio prediction of nuclear electric quadrupole observables by scaling to the charge radius. Phys. Rev. C 2022, 105, L061302. [Google Scholar] [CrossRef]
  49. Ewald, G.; Nörtershäuser, W.; Dax, A.; Götte, S.; Kirchner, R.; Kluge, H.-J.; Kühl, T.; Sanchez, R.; Wojtaszek, A.; Bushaw, B.A.; et al. Nuclear charge radii of 8,9Li determined by laser spectroscopy. Phys. Rev. Lett. 2004, 93, 113002. [Google Scholar] [CrossRef]
  50. Sánchez, R.; Nörtershäuser, W.; Ewald, G.; Albers, D.; Behr, J.; Bricault, P.; Bushaw, B.A.; Dax, A.; Dilling, J.; Dombsky, M.; et al. Nuclear charge radii of 9,11Li: The influence of halo neutrons. Phys. Rev. Lett. 2006, 96, 033002. [Google Scholar] [CrossRef]
  51. Nörtershäuser, W.; Tiedemann, D.; Žáková, M.; Andjelkovic, Z.; Blaum, K.; Bissell, M.L.; Cazan, R.; Drake, G.W.F.; Geppert, C.; Kowalska, M.; et al. Nuclear charge radii of 7,9,10Be and the one-neutron halo nucleus 11Be. Phys. Rev. Lett. 2009, 102, 062503. [Google Scholar] [CrossRef]
  52. Brown, B.A. Mirror charge radii and the neutron equation of state. Phys. Rev. Lett. 2017, 119, 122502. [Google Scholar] [CrossRef]
  53. Brown, B.A.; Minamisono, K.; Piekarewicz, J.; Hergert, H.; Garand, D.; Klose, A.; König, K.; Lantis, J.D.; Liu, Y.; Maaß, B.; et al. Implications of the 36Ca36S and 38Ca38Ar difference in mirror charge radii on the neutron matter equation of state. Phys. Rev. Res. 2020, 2, 022035. [Google Scholar] [CrossRef]
  54. Pineda, S.V.; König, K.; Rossi, D.M.; Brown, B.A.; Incorvati, A.; Lantis, J.; Minamisono, K.; Nörtershäuser, W.; Piekarewicz, J.; Powel, R.; et al. Charge radius of neutron-deficient 54Ni and symmetry energy constraints using the difference in mirror pair charge radii. Phys. Rev. Lett. 2021, 127, 182503. [Google Scholar] [CrossRef] [PubMed]
  55. Naito, T.; Roca-Maza, X.; Colò, G.; Liang, H.; Sagawa, H. Isospin symmetry breaking in the charge radius difference of mirror nuclei. Phys. Rev. C 2022, 106, L061306. [Google Scholar] [CrossRef]
  56. Reinhard, P.G.; Nazarewicz, W. Information content of the differences in the charge radii of mirror nuclei. Phys. Rev. C 2022, 105, L021301. [Google Scholar] [CrossRef]
  57. Bano, P.; Pattnaik, S.P.; Centelles, M.; Viñas, X.; Routray, T.R. Correlations between charge radii differences of mirror nuclei and stellar observables. Phys. Rev. C 2023, 108, 015802. [Google Scholar] [CrossRef]
  58. Yang, J.; Piekarewicz, J. Difference in proton radii of mirror nuclei as a possible surrogate for the neutron skin. Phys. Rev. C 2018, 97, 014314. [Google Scholar] [CrossRef]
  59. Gaidarov, M.; Moumene, I.; Antonov, A.; Kadrev, D.; Sarriguren, P.; Moya de Guerra, E. Proton and neutron skins and symmetry energy of mirror nuclei. Nucl. Phys. A 2020, 1004, 122061. [Google Scholar] [CrossRef]
  60. Novario, S.J.; Lonardoni, D.; Gandolfi, S.; Hagen, G. Trends of neutron skins and radii of mirror nuclei from first principles. Phys. Rev. Lett. 2023, 130, 032501. [Google Scholar] [CrossRef]
  61. Maaß, B.; Müller, P.; Nörtershäuser, W.; Clark, J.; Gorges, C.; Kaufmann, S.; König, K.; Krämer, J.; Levand, A.; Orford, R.; et al. Towards laser spectroscopy of the proton-halo candidate boron-8. Hyperfine Interact. 2017, 238, 25. [Google Scholar] [CrossRef]
  62. Yerokhin, V.A.; Pachucki, K. Theoretical energies of low-lying states of light helium-like ions. Phys. Rev. A 2010, 81, 022507. [Google Scholar] [CrossRef]
  63. Yerokhin, V.A.; Patkóš, V.; Pachucki, K. QED calculations of energy levels of heliumlike ions with 5 ≤ Z ≤ 30. Phys. Rev. A 2022, 106, 022815. [Google Scholar] [CrossRef]
  64. Yerokhin, V.A.; Patkóš, V.; Pachucki, K. QED 7 effects for triplet states of heliumlike ions. Phys. Rev. A 2023, 107, 012810. [Google Scholar] [CrossRef]
  65. Riis, E.; Sinclair, A.G.; Poulsen, O.; Drake, G.W.F.; Rowley, W.R.C.; Levick, A.P. Lamb shifts and hyperfine structure in 6Li+ and 7Li+: Theory and experiment. Phys. Rev. A 1994, 49, 207–220. [Google Scholar] [CrossRef]
  66. Guan, H.; Chen, S.; Qi, X.-Q.; Liang, S.; Sun, W.; Zhou, P.; Huang, Y.; Zhang, P.-P.; Zhong, Z.-X.; Yan, Z.-C.; et al. Probing atomic and nuclear properties with precision spectroscopy of fine and hyperfine structures in the 7Li+ ion. Phys. Rev. A 2020, 102, 030801. [Google Scholar] [CrossRef]
  67. Clausen, G.; Jansen, P.; Scheidegger, S.; Agner, J.A.; Schmutz, H.; Merkt, F. Ionization energy of the metastable 2 1S0 state of 4He from Rydberg-series extrapolation. Phys. Rev. Lett. 2021, 127, 093001. [Google Scholar] [CrossRef] [PubMed]
  68. Clausen, G.; Scheidegger, S.; Agner, J.A.; Schmutz, H.; Merkt, F. Imaging-assisted single-photon Doppler-free laser spectroscopy and the ionization energy of metastable triplet helium. Phys. Rev. Lett. 2023, 131, 103001. [Google Scholar] [CrossRef] [PubMed]
  69. Maaß, B. Laser Spectroscopy of the Boron Isotopic Chain. Ph.D. Thesis, TU Darmstadt, Darmstadt, Germany, 2020. [Google Scholar] [CrossRef]
  70. König, K.; Krämer, K.; Geppert, C.; Imgram, P.; Maaß, B.; Ratajczyk, T.; Nörtershäuser, W. A new Collinear Apparatus for Laser Spectroscopy and Applied Science (COALA). Rev. Sci. Instrum. 2020, 91, 081301. [Google Scholar] [CrossRef] [PubMed]
  71. Mohr, K.; Buß, A.; Andelkovic, Z.; Hannen, V.; Horst, M.; Imgram, P.; König, K.; Maaß, B.; Nörtershäuser, W.; Rausch, S.; et al. Collinear laser spectroscopy of helium-like 11B3+. Atoms 2023, 11, 11. [Google Scholar] [CrossRef]
  72. Imgram, P. High-Precision Laser Spectroscopy of Helium-like Carbon 12C4+. Ph.D. Thesis, TU Darmstadt, Darmstadt, Germany, 2023. [Google Scholar] [CrossRef]
  73. Imgram, P.; König, K.; Maaß, B.; Müller, P.; Nörtershäuser, W. Collinear laser spectroscopy of highly charged ions produced with an electron-beam ion source. Phys. Rev. A 2023, 108, 062809. [Google Scholar] [CrossRef]
  74. Imgram, P.; König, K.; Maaß, B.; Müller, P.; Nörtershäuser, W. Collinear laser spectroscopy of 2 3S1 → 2 3PJ transitions in helium-like 12C4+. Phys. Rev. Lett. 2023, 131, 243001. [Google Scholar] [CrossRef] [PubMed]
  75. Sailer, T.; Debierre, V.; Harman, Z.; Heiße, F.; König, C.; Morgner, J.; Tu, B.; Volotka, A.V.; Keitel, C.H.; Blaum, K.; et al. Measurement of the bound-electron g-factor difference in coupled ions. Nature 2022, 606, 479–483. [Google Scholar] [CrossRef]
  76. Schmidt, S.; Willig, M.; Haack, J.; Horn, R.; Adamczak, A.; Abdou Ahmed, M.; Amaro, F.D.; Amaro, P.; Biraben, F.; Carvalho, P.; et al. The next generation of laser spectroscopy experiments using light muonic atoms. J. Phys. Conf. Ser. 2018, 1138, 012010. [Google Scholar] [CrossRef]
  77. Yerokhin, V.A. Hyperfine structure of Li and Be+. Phys. Rev. A 2008, 78, 012513. [Google Scholar] [CrossRef]
  78. Sun, W.; Zhang, P.-P.; Zhou, P.-P.; Chen, S.-L.; Zhou, Z.-Q.; Huang, Y.; Qi, X.-Q.; Yan, Z.-C.; Shi, T.-Y.; Drake, G.W.F.; et al. Measurement of hyperfine structure and the Zemach radius in 6Li+ using optical Ramsey technique. Phys. Rev. Lett. 2023, 131, 103002. [Google Scholar] [CrossRef]
  79. Pachucki, K.; Patkóš, V.; Yerokhin, V.A. Hyperfine splitting in 6,7Li+. Phys. Rev. A 2023, 108, 052802. [Google Scholar] [CrossRef]
  80. Pachucki, K.; Patkóš, V.; Yerokhin, V.A. Accurate determination of 6,7Li nuclear magnetic moments. Phys. Lett. B 2023, 846, 138189. [Google Scholar] [CrossRef]
  81. Krutov, A.A.; Martynenko, A.P.; Martynenko, F.A.; Sukhorukova, O.S. Lamb shift in muonic ions of lithium, beryllium, and boron. Phys. Rev. A 2016, 94, 062505. [Google Scholar] [CrossRef]
  82. Dorokhov, A.E.; Faustov, R.N.; Martynenko, A.P.; Martynenko, F.A. Precision physics of muonic ions of lithium, beryllium and boron. Int. J. Mod. Phys. A 2021, 36, 2150022. [Google Scholar] [CrossRef]
  83. Li Muli, S.S.; Poggialini, A.; Bacca, S. Muonic lithium atoms: Nuclear structure corrections to the Lamb shift. SciPost Phys. Proc. 2020, 3, 028. [Google Scholar] [CrossRef]
  84. Valuev, I.A.; Colò, G.; Roca-Maza, X.; Keitel, C.H.; Oreshkina, N.S. Evidence against nuclear polarization as source of fine-structure anomalies in muonic atoms. Phys. Rev. Lett. 2022, 128, 203001. [Google Scholar] [CrossRef]
  85. Barrett, B.R.; Navrátil, P.; Vary, J.P. Ab initio no core shell model. Prog. Part. Nucl. Phys. 2013, 69, 131–181. [Google Scholar] [CrossRef]
  86. Lanczos, C. An iteration method for the solution of the eigenvalue problem of linear differential and integral operators. J. Res. Natl. Bur. Stand. 1950, 45, 255–282. [Google Scholar] [CrossRef]
  87. Haydock, R. The inverse of a linear operator. J. Phys. Math. Nucl. Gen. 1974, 7, 2120–2124. [Google Scholar] [CrossRef]
  88. Reuter, W.; Fricke, G.; Merle, K.; Miska, H. Nuclear charge distribution and rms radius of 12C from absolute elastic electron scattering measurements. Phys. Rev. C 1982, 26, 806–818. [Google Scholar] [CrossRef]
  89. Schneuwly, H.; Vogel, P. Electronic K X-ray energies in heavy muonic atoms. Phys. Rev. A 1980, 22, 2081–2087. [Google Scholar] [CrossRef]
  90. Simons, L.M.; Abbot, D.; Bach, B.; Bacher, R.; Badertscher, A.; Blüm, P.; DeCecco, P.; Eades, J.; Egger, J.; Elsener, K.; et al. Exotic atoms and their electron shell. Nucl. Instrum. Meth. Phys. Res. B Beam Interact. Mater. Atoms 1994, 87, 293–300. [Google Scholar] [CrossRef]
  91. Desclaux, J.P. A multiconfiguration relativistic DIRAC-FOCK program. Comput. Phys. Commun. 1975, 9, 31–45. [Google Scholar] [CrossRef]
  92. Mallow, J.V.; Desclaux, J.P.; Freeman, A.J. Dirac-Fock method for muonic atoms: Transitions energies, wave functions, and charge densities. Phys. Rev. A 1978, 17, 1804–1809. [Google Scholar] [CrossRef]
  93. Santos, J.P.; Parente, F.; Boucard, S.; Indelicato, P.; Desclaux, J.P. X-ray energies of circular transitions and electron screening in kaonic atoms. Phys. Rev. A 2005, 71, 032501. [Google Scholar] [CrossRef]
  94. Trassinelli, M.; Indelicato, P. Relativistic calculations of pionic and kaonic atoms’ hyperfine structure. Phys. Rev. A 2007, 76, 012510. [Google Scholar] [CrossRef]
  95. Indelicato, P. Nonperturbative evaluation of some QED contributions to the muonic hydrogen n = 2 Lamb shift and hyperfine structure. Phys. Rev. A 2013, 87, 022501. [Google Scholar] [CrossRef]
  96. De Wit, S.A.; Backenstoss, G.; Daum, C.; Sens, J.C.; Acker, H.L. Measurement and analysis of muonic X-ray spectra in deformed nuclei. Nucl. Phys. 1966, 87, 657–702. [Google Scholar] [CrossRef]
  97. Pieper, W.; Greiner, W. The influence of nuclear dynamics on the X-ray spectrum of muonic atoms. Phys. Lett. B 1967, 24, 377–380. [Google Scholar] [CrossRef]
  98. Foot, C.J.; Stacey, D.N.; Stacey, V.; Kloch, R.; Leś, Z. Isotope effects in the nuclear charge distribution in zinc. Proc. R. Soc. A Math. Phys. Engin. Sci. 1982, 384, 205–216. [Google Scholar] [CrossRef]
  99. Sahoo, B.K.; Ohayon, B. All-optical differential radii in zinc. Phys. Rev. Res. 2023, 5, 043142. [Google Scholar] [CrossRef]
  100. Çavuşoǧlu, A.; Sikora, B. Impact of the nuclear charge distribution on the g-factors and ground state energies of bound muons. arXiv 2023, arXiv:2311.16855. [Google Scholar] [CrossRef]
  101. Paul, N.; Bian, G.; Azuma, T.; Okada, S.; Indelicato, P. Testing quantum electrodynamics with exotic atoms. Phys. Rev. Lett. 2021, 126, 173001. [Google Scholar] [CrossRef] [PubMed]
  102. Indelicato, P. QED tests with highly charged ions. J. Phys. B At. Mol. Opt. Phys. 2019, 52, 232001. [Google Scholar] [CrossRef]
  103. Fricke, G.; Heilig, K. Nuclear Charge Radii; Springer Nature Switzerland AG: Cham, Switzerland, 2004. [Google Scholar] [CrossRef]
  104. Simons, L.M. X-ray spectroscopy at PSI. Hyperfine Interact. 1999, 119, 281–290. [Google Scholar] [CrossRef]
  105. Ruckstuhl, W.; Aas, B.; Beer, W.; Beltrami, I.; de Boer, F.W.N.; Bos, K.; Goudsmit, P.F.A.; Kiebele, U.; Leisi, H.J.; Strassner, G.; et al. High-precision muonic X-ray measurement of the rms charge radius of 12C with a crystal spectrometer. Phys. Rev. Lett. 1982, 49, 859–862. [Google Scholar] [CrossRef]
  106. Fleischmann, A.; Enss, C.; Seidel, G. Metallic magnetic calorimeters. In Cryogenic Particle Detection; Enss, C., Ed.; Springer: Berlin/Heidelberg, Germany, 2005; pp. 151–216. [Google Scholar] [CrossRef]
  107. Unger, D.; Abeln, A.; Cocolios, T.E.; Eizenberg, O.; Enss, C.; Fleischmann, A.; Gastaldo, L.; Godinho, C.; Heines, M.; Hengstler, D.; et al. MMC array to study X-ray transitions in muonic atoms. arXiv 2023, arXiv:2311.12014. [Google Scholar] [CrossRef]
  108. Hengstler, D.; Keller, M.; Schöltz, C.; Geist, J.; Krantz, M.; Kempf, S.; Gastaldo, L.; Fleischmann, A.; Gassner, T.; Weber, G.; et al. Towards FAIR: First measurements of metallic magnetic calorimeters for high-resolution X-ray spectroscopy at GSI. Phys. Scr. 2015, T166, 014054. [Google Scholar] [CrossRef]
  109. Unger, D.; Abeln, A.; Enss, C.; Fleischmann, A.; Hengstler, D.; Kempf, S.; Gastaldo, L. High-resolution for IAXO: MMC-based X-ray detectors. J. Instrum. 2021, 16, P06006. [Google Scholar] [CrossRef]
  110. Wauters, F.; Knecht, A. The muX project. SciPost Phys. Proc. 2021, 5, 022. [Google Scholar] [CrossRef]
  111. Boyd, S.; Kim, G.B.; Hall, J.; Cantor, R.; Friedrich, S. Metallic magnetic calorimeters for high-accuracy nuclear decay data. J. Low Temp. Phys. 2020, 199, 681–687. [Google Scholar] [CrossRef]
  112. Deslattes, R.D.; Kessler, E.G.; Indelicato, P.; de Billy, L.; Lindroth, E.; Anton, J. X-ray transition energies: New approach to a comprehensive evaluation. Rev. Mod. Phys. 2003, 75, 35–99. [Google Scholar] [CrossRef]
  113. Mendenhall, M.H.; Hudson, L.T.; Szabo, C.I.; Henins, A.; Cline, J.P. The molybdenum K-shell X-ray emission spectrum. J. Phys. At. Mol. Opt. Phys. 2019, 52, 215004. [Google Scholar] [CrossRef] [PubMed]
  114. Hudson, L.T.; Cline, J.P.; Henins, A.; Mendenhall, M.H.; Szabo, C. Contemporary X-ray wavelength metrology and traceability. Radiat. Phys. Chem. 2020, 167, 108392. [Google Scholar] [CrossRef] [PubMed]
  115. Trassinelli, M.; Anagnostopoulos, D.F.; Borchert, G.; Dax, A.; Egger, J.-P.; Gotta, D.; Hennebach, M.; Indelicato, P.; Liu, Y.-W.; Manil, B.; et al. Measurement of the charged pion mass using X-ray spectroscopy of exotic atoms. Phys. Lett. B 2016, 759, 583–588. [Google Scholar] [CrossRef]
  116. Okumura, T.; Azuma, T.; Bennett, D.A.; Chiu, I.; Doriese, W.B.; Durkin, M.S.; Fowler, J.W.; Gard, J.D.; Hashimoto, T.; Hayakawa, R.; et al. Proof-of-principle experiment for testing strong-field quantum electrodynamics with exotic atoms: High precision X-ray spectroscopy of muonic neon. Phys. Rev. Lett. 2023, 130, 173001. [Google Scholar] [CrossRef] [PubMed]
  117. Curceanu, C.; Abbene, L.; Amsler, C.; Bazzi, M.; Bettelli, M.; Borghi, G.; Bosnar, D.; Bragadireanu, M.; Buttacavoli, A.; Cargnelli, M.; et al. Kaonic atoms at the DAΦNE collider: A strangeness adventure. Front. Phys. 2023, 11, 1240250. [Google Scholar] [CrossRef]
Table 1. Current status of the charge radii, r c , of light stable nuclei. The method abbreviations are as follows: ‘ μ -Laser’—muonic atoms laser spectroscopy; ‘el. scat.’—elastic electron scattering; ‘OIS’—optical isotope shift; ‘ π + scat.’—elastic positive pion scattering; ‘ μ -X’—muonic atom X-ray spectroscopy. Numbers in the parentheses mark the uncertainty to the preceding digits, which is also denoted by σ r .
Table 1. Current status of the charge radii, r c , of light stable nuclei. The method abbreviations are as follows: ‘ μ -Laser’—muonic atoms laser spectroscopy; ‘el. scat.’—elastic electron scattering; ‘OIS’—optical isotope shift; ‘ π + scat.’—elastic positive pion scattering; ‘ μ -X’—muonic atom X-ray spectroscopy. Numbers in the parentheses mark the uncertainty to the preceding digits, which is also denoted by σ r .
r c , fm σ r r c 1 / 10 3 MethodReferences
1H0.84060(39)0.5 μ -Laser[27,28]
2H 2.12775 ( 17 ) 0.1OIS + r c (61H)[29,30]
3He 1.97007 ( 94 ) 0.5 μ -Laser[28,31]
4He 1.6786 ( 12 )    0.7 μ -Laser[28,32]
6Li 2.589 ( 39 )      15  el. scat.[33,34,35,36]
7Li 2.444 ( 42 )      17  OIS + r c (6Li)[36,37]
9Be 2.519 ( 32 )      13  el. scat. 1[38]
10B 2.510 ( 31 )      12  OIS+ r c (11B)[39]
11B 2.411 ( 21 )      8.7 π + scat. + r c (12C)[40]
12C 2.4829 ( 19 )    0.8 μ -X[41]
13C 2.4628 ( 39 )    1.6 μ -X[42]
1 Estimated using a model-dependent analysis of an electron scattering experiment covering a narrow momentum transfer range. The same study quotes a radius for 12C [38], which differs by three standard deviations from the radius modern value. A systematic uncertainty to this value was added in quadrature to the smaller experimental uncertainty.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ohayon, B.; Abeln, A.; Bara, S.; Cocolios, T.E.; Eizenberg, O.; Fleischmann, A.; Gastaldo, L.; Godinho, C.; Heines, M.; Hengstler, D.; et al. Towards Precision Muonic X-ray Measurements of Charge Radii of Light Nuclei. Physics 2024, 6, 206-215. https://doi.org/10.3390/physics6010015

AMA Style

Ohayon B, Abeln A, Bara S, Cocolios TE, Eizenberg O, Fleischmann A, Gastaldo L, Godinho C, Heines M, Hengstler D, et al. Towards Precision Muonic X-ray Measurements of Charge Radii of Light Nuclei. Physics. 2024; 6(1):206-215. https://doi.org/10.3390/physics6010015

Chicago/Turabian Style

Ohayon, Ben, Andreas Abeln, Silvia Bara, Thomas Elias Cocolios, Ofir Eizenberg, Andreas Fleischmann, Loredana Gastaldo, César Godinho, Michael Heines, Daniel Hengstler, and et al. 2024. "Towards Precision Muonic X-ray Measurements of Charge Radii of Light Nuclei" Physics 6, no. 1: 206-215. https://doi.org/10.3390/physics6010015

Article Metrics

Back to TopTop